...

Model Abstraction Techniques for Soil-Water Flow and Transport

by user

on
Category: Documents
21

views

Report

Comments

Transcript

Model Abstraction Techniques for Soil-Water Flow and Transport
NUREG/CR-6884
Model Abstraction Techniques
for Soil-Water Flow and
Transport
U.S. Department of Agriculture
Agricultural Research Service
U.S. Nuclear Regulatory Commission
Office of Nuclear Regulatory Research
Washington, DC 20555-0001
AVAILABILITY OF REFERENCE MATERIALS
IN NRC PUBLICATIONS
NRC Reference Material
Non-NRC Reference Material
As of November 1999, you may electronically access
NUREG-series publications and other NRC records at
NRC’s Public Electronic Reading Room at
http://www.nrc.gov/reading-rm.html. Publicly released
records include, to name a few, NUREG-series
publications; Federal Register notices; applicant,
licensee, and vendor documents and correspondence;
NRC correspondence and internal memoranda;
bulletins and information notices; inspection and
investigative reports; licensee event reports; and
Commission papers and their attachments.
Documents available from public and special technical
libraries include all open literature items, such as
books, journal articles, and transactions, Federal
Register notices, Federal and State legislation, and
congressional reports. Such documents as theses,
dissertations, foreign reports and translations, and
non-NRC conference proceedings may be purchased
from their sponsoring organization.
NRC publications in the NUREG series, NRC
regulations, and Title 10, Energy, in the Code of
Federal Regulations may also be purchased from one
of these two sources.
1. The Superintendent of Documents
U.S. Government Printing Office
Mail Stop SSOP
Washington, DC 20402–0001
Internet: bookstore.gpo.gov
Telephone: 202-512-1800
Fax: 202-512-2250
2. The National Technical Information Service
Springfield, VA 22161–0002
www.ntis.gov
1–800–553–6847 or, locally, 703–605–6000
A single copy of each NRC draft report for comment is
available free, to the extent of supply, upon written
request as follows:
Address: Office of Administration,
Reproduction and Distribution
Services Section,
U.S. Nuclear Regulatory Commission
Washington, DC 20555-0001
E-mail:
[email protected]
Facsimile: 301–415–2289
Some publications in the NUREG series that are
posted at NRC’s Web site address
http://www.nrc.gov/reading-rm/doc-collections/nuregs
are updated periodically and may differ from the last
printed version. Although references to material found
on a Web site bear the date the material was accessed,
the material available on the date cited may
subsequently be removed from the site.
Copies of industry codes and standards used in a
substantive manner in the NRC regulatory process are
maintained at—
The NRC Technical Library
Two White Flint North
11545 Rockville Pike
Rockville, MD 20852–2738
These standards are available in the library for
reference use by the public. Codes and standards are
usually copyrighted and may be purchased from the
originating organization or, if they are American
National Standards, from—
American National Standards Institute
11 West 42nd Street
New York, NY 10036–8002
www.ansi.org
212–642–4900
Legally binding regulatory requirements are stated
only in laws; NRC regulations; licenses, including
technical specifications; or orders, not in
NUREG-series publications. The views expressed
in contractor-prepared publications in this series are
not necessarily those of the NRC.
The NUREG series comprises (1) technical and
administrative reports and books prepared by the
staff (NUREG–XXXX) or agency contractors
(NUREG/CR–XXXX), (2) proceedings of
conferences (NUREG/CP–XXXX), (3) reports
resulting from international agreements
(NUREG/IA–XXXX), (4) brochures
(NUREG/BR–XXXX), and (5) compilations of legal
decisions and orders of the Commission and Atomic
and Safety Licensing Boards and of Directors’
decisions under Section 2.206 of NRC’s regulations
(NUREG–0750).
DISCLAIMER: This report was prepared as an account of work sponsored by an agency of the U.S. Government.
Neither the U.S. Government nor any agency thereof, nor any employee, makes any warranty, expressed or
implied, or assumes any legal liability or responsibility for any third party’s use, or the results of such use, of any
information, apparatus, product, or process disclosed in this publication, or represents that its use by such third
party would not infringe privately owned rights.
NUREG/CR-6884
Model Abstraction Techniques
for Soil-Water Flow and
Transport
Manuscript Completed: September 2005
Date Published: December 2006
Prepared by
Y.A. Pachepsky1, A.K. Guber 3, M.T. Van Genuchten 2,
T.J. Nicholson 4, R.E. Cady 4, J. Simunek 3, M.G. Schaap 3
1
U.S. Department of Agriculture
Agricultural Research Service
Environmental Microbial Safety Laboratory
Beltsville, MD 20705
3
University of California Riverside
Department of Environmental Sciences
Riverside, CA 92521
T.J. Nicholson, NRC Project Manager
Prepared for
Division of Fuel, Engineering & Radiological Research
Office of Nuclear Regulatory Research
U.S. Nuclear Regulatory Commission
Washington, DC 20555-0001
NRC Job Code Y6724
2
U.S. Department of Agriculture
Agricultural Research Service
George E. Brown Jr. Salinity Laboratory
Riverside, CA 92705
4
U.S. Nuclear Regulatory Commission
Office of Nuclear Regulatory Research
Washington, DC 20555-0001
ABSTRACT
This report describes the methodology of model abstraction in subsurface hydrology.
Model abstraction is defined as the methodology for reducing the complexity of a
simulation model while maintaining the validity of the simulation results with respect to
the question that the simulation is being used to address. The need in model abstraction
may stem from the need to improve the reliability and reduce uncertainty of simulations,
to make the modeling and its results more explicable and transparent to the end users, and
to enable more efficient use of available resources in data collection and computations. A
comprehensive review of model simplification techniques developed in subsurface
hydrology is included. Abstractions of both model structure and model parameter
determination are described. A systematic and objective approach to model abstraction is
outlined. A case study is presented that is designed to illustrate how model abstraction
can affect performance assessment of contaminant migration at a relatively humid site.
Although the model abstraction methodology is generic, it is designed to be of practical
use to NRC licensing staff in their review of the performance assessment of
decommissioning sites and waste disposal facilities. The model abstraction process
provides a systematic approach to understand the adequacy of model simplification, and
facilitates communication and transparency of the model to regulators, stakeholders, and
the general public.
iii
FOREWORD
An important step in the U.S. Nuclear Regulatory Commission (NRC) staff evaluation of the
proposed decommissioning of nuclear sites and facilities is the review of licensees' analyses of the
performance of the proposed actions during extended periods of future time. In complex cases,
licensees often must develop their own conceptual models of their sites and facilities, and use a
variety of computer codes to model the potential radionuclide transport pathways. The model
development process is often not systematic and does not identify simplifications made to reduce the
complexity of the site’s features, events, and processes (FEPs) or explain that the simulation results
honor the significant FEPs affecting subsurface flow and radionuclide transport assessments. To
better understand the model development process and its implications on modeling of complex
cases, the NRC partnered with the U.S. Department of Agriculture’s (USDA) Agricultural Research
Service (ARS) under Interagency Agreement (RES-02-008), “Model Abstraction Techniques for
Soil Water Flow and Transport.” The objective was to identify, test, and confirm the practicality of
various model abstraction techniques for establishing the appropriate combination of site specific
models and supporting data collection programs to simulate water flow and contaminant transport
through soils.
This report documents the conduct and findings of this research. The research findings provide the
technical basis for developing guidance to NRC licensees and NRC licensing staff on reviewing
models of water flow and solute transport in soils, sediments, and porous rocks. The report presents
a systematic approach for identifying the need for model abstractions, identifies and reviews model
abstraction techniques, discusses the process of selection of appropriate model abstraction
techniques, and demonstrates the testing and confirmation of the model abstraction approach through
an application to field data. The case study demonstrates the effectiveness of model abstraction in
selecting an appropriate level of simplicity for simulating soil water fluxes at a relatively humid site
where transport may be affected by the presence of soil macropores and related preferential flow
phenomena. The case study shows the capabilities, limitations, and usefulness of selected model
abstraction methods and confirms the utility and correctness of the model abstraction method
through a series of comparisons. The model abstraction process addresses the uncertainty in model
structure and in model parameter determination, and facilitates transparency and realism in the sitespecific modeling of essential FEPs. The results of this research supports regulatory decisions that
will help make modeling activities and decisions on which they are based more efficient and
effective.
This report is a technology transfer document. The NRC staff can use it in reviewing subsurface
hydrologic models. The implementation of the model abstraction methodology provides a context
for more efficient use of modeling in the regulatory process. Earlier versions of this report were
used in training workshops at NRC headquarters on June 22─24, 2004, and May 31, 2005. The
report was also circulated to cooperating Federal agencies working under the memorandum of
understanding on research in multimedia environmental models. The report was peer-reviewed by
the researchers who earlier developed the general strategy of hydrologic modeling and uncertainty
analysis for nuclear sites and facilities.
Carl J. Paperiello, Director
Office of Nuclear Regulatory Research
v
CONTENTS
ABSTRACT..............................................................................................................
FOREWORD.....................................................................................................…....
EXECUTIVE SUMMARY ......................................................................................
ACKNOWLEDGEMENTS...............................................................................…...
ABBREVIATIONS.........................................................................................……..
SYMBOLS…............….....…............................................................................……
1 INTRODUCTION………………………………………………………………
1.1 Background…………………………………………………………………..
1.2 Model abstraction……………….………………………………………….
1.3 Objectives…………………………………………………………….………
2 MODEL ABSTRACTION TECHNIQUES………………………………………
2.1 Background…………………………………………………………………….
2.2 Model abstraction techniques for flow and transport models……………...
2.2.1 Hierarchies of models…………………………………………………….
2.2.2 Delimited input domain…………………………………………………..
2.2.3 Scale change………………………………………………………………
2.2.3.1 Scale change – upscaling…………………………………………
2.2.3.2 Scale change – aggregation……………………………………….
2.2.4 Metamodeling……………………………………………………………
2.2.5 Discretization……………………………………………………………..
2.2.6 Scaling…………………………………………………………………….
2.2.7 Parameter estimation with pedotransfer functions………………………..
2.2.7.1. Estimating water flow parameters………………………………..
2.2.7.2. Performance of pedotransfer functions…………………………..
2.2.7.3 Emerging pedotransfer techniques………………………………..
2.3. Model abstraction implementation…………………………………………...
2.3.1 Status of model abstraction in flow and transport modeling……………...
2.3.2. Model abstraction process………………………………………………...
2.3.2.1 Justify the need for model abstraction……………………………
2.3.2.1.1 Need in abstraction because of the limited observation
ability………………………………………………….
2.3.2.1.2 Need in abstraction because of uncertainty in key model
outputs…………………………………………………
2.3.2.2 Review the context of the modeling problem……………………..
2.3.2.3 Define applicable MA techniques…………………………………
2.3.2.4 Define and execute model abstractions……………………………
2.4 Concluding remarks…………………………………………………………...
3 MODEL ABSTRACTION CASE STUDY……………………………………….
3.1 Introduction…………………………………………………………………….
3.2 Description of experiments…………………………………………………….
3.2.1 Field setup…………………………………………………………………
3.2.2 Dataset overview………………………………………………………….
3.2.2.1 Precipitation……………………………………………………….
3.2.2.2 Soil water contents, capillary pressures and soil temperatures……
vii
iii
v
xiii
xvi
xx
xx
1
1
2
3
5
5
11
11
14
16
16
19
19
22
22
22
26
33
36
38
38
39
39
40
43
44
45
47
48
51
51
51
51
59
59
59
3.2.2.3 Soil water fluxes…………………………………………………...
3.2.2.4 Laboratory data on soil water retention…………………………..
3.3 The base model and the need in model abstraction………………………….
3.4 Review of the model abstraction context……………………………………..
3.4.1.The key output…………………………………………………………….
3.4.2 Statistics to evaluate model abstractions………………………………….
3.4.3 Available data and their uncertainty………..……………………………..
3.4.3.1 Averaging water contents …...…………………………………….
3.4.3.2 Temporal persistence in soil water contents...…………………….
3.4.3.3 Surface evaporation……………………………………………….
3.4.4 Additional information to ensure sound abstraction……………………...
3.4.4.1 Evidence of the macropore flow………………………………….
3.4.5 Recalibration of the base model………………………………………….
3.5 Selection and application of model abstraction techniques………………..
3.5.1 Model abstraction design………………………………………………..
3.5.2 Abstraction with a hierarchy of models: using water budget……………
3.5.3 Abstraction with the scale change: profile aggregation…………………
3.5.4 Abstraction of parameter determination: using an ensemble of
pedotransfer functions…………………………………………………...
3.5.5 Sequential model abstractions…………………………………………..
3.5.6 Abstraction with metamodeling………………………………………...
3.5.7 Pedotransfer-estimated vs. laboratory-measured water retention………
3.6 Summary and conclusion…………………………………………………….
4 REFERENCES………...…………………………………………………………...
APPENDICES
A. PEDOTRANSFER FUNCTIONS TO ESTIMATE SOIL WATER
RETENTION…………………………………………………………………………
B. DIURNAL OSCILLATIONS IN WATER CONENT AND CAPILLARY
PRESSURE DATA…………………………………………………………………..
C. USING TEMPORAL PERSISTENSE TO ESTIMATE DEPTHAVERAGED WATER CONTENTS WITH MISSING DATA…………………...
D. ESTIMATING MATRIC UNSATURATED HYDRAULIC
CONDUCTIVITY IN DEEP SOIL LAYERS……………………………………...
E. MWBUS: ONE-DIMENSIONAL SOIL WATER FLOW MODEL…………..
viii
64
64
64
73
73
73
75
75
75
83
83
83
88
92
92
92
93
99
99
110
116
118
121
A-1
B-1
C-1
D-1
E-1
Figures
2-1
2-2
2-3
2-4
2-5
2-6
2-7
2-8
2-9
2-10
2-11
2-12
3-1
3-2
3-3
3-4
3-5
3-6
3-7
3-8
3-9
3-10
3-11
3-12
3-13
3-14
3-15
3-16
General taxonomy of model abstraction techniques …………………...….…
Categories of model abstraction techniques relevant to flow and transport
modeling in subsurface hydrology……………………………………………
Hierarchy of models to simulate water flow and solute transport in structured
soils or in unsaturated fractured rock…………………………………………
Maximum tracer velocity in 34 documented field experiments with variably
saturated soils…………………………………………………………………
Spatial and temporal operational scales in hydrology………………………
Soil field capacity and soil water retention at fixed matric potentials………
Observed dependencies of the solute dispersivity on distance in unsaturated
soils…………………………………………………………………………..
Examples of scale-dependence in soil saturated hydraulic conductivity ……
Relationship between field and laboratory water contents at the same soilwater matric potential…………………………………………………………
Saturated hydraulic conductivity Ksat grouped by textural classes and bulk
density classes; (a) median values Ksat(50%) for the groups, (b) difference
between third Ksat(75%) and first Ksat(25%) quartiles of the distributions within
each group; data for high and low bulk density samples are shown by filled
and hollow symbols, respectively……………………………………………..
Estimating water contents at 33 kPa (○) and at 1500 kPa (●) with 21 PTFs
developed in various region of the world……………………………………
Using the dense auxiliary data to infer soil hydraulic properties……………
Soil profile at the Bekkevoort experimental site……………………...………
View of the instrumented trench……………………………………………
Locations and numbering of TDR-probes (z), tensiometers ({) and
temperature probes (•)……………………………………………………….
Three passive capillary samplers on the PVC plate…………………………
Installation of passive capillary samplers……………………………………
Cumulative rainfall and daily precipitation during the experiment…………
Time series of average water contents at each of the measurement depths…
Time series of average capillary pressures at each of the measurement depths
Consecutive images of two-dimensional distributions of soil moisture
contents; days from the beginning of the experiment are shown……………
Time series of soil temperature……………………………………………….
Cumulative rainfall and soil water fluxes measured with passive capillary
lysimeters (PCAPS) at two depths……………………………………………
Van Genuchten parameters derived from water retention of samples taken
from three soil horizons along the 30-m trench……………………………….
Measured cumulative rainfall and simulated cumulative actual infiltration for
the different objective functions………………………………………………
Wetting-drying periods selected to compute the cumulative water flux as the
key output of the base and abstracted models………………………………...
Time series of TDR-measured water contents at the depth of 15 cm (a-d) and
precipitation (e) during the period when all probes worked at this depth…….
Cumulative change in soil water storage in several locations along the
ix
7
12
13
15
17
20
23
24
25
30
35
37
52
54
55
57
58
61
62
63
66
67
68
69
72
74
77
3-17
3-18
3-19
3-20
3-21
3-22
3-23
3-24
3-25
3-26
3-27
3-28
3-29
3-30
3-31
3-32
3-33
3-34
3-35
3-36
3-37
3-38
trench. Location numbering is given in the Fig. 3-3…………………………
Capillary pressure gradient during a drying period in the location 10………..
Daily average water content and pressure head time series at the location 10..
Water contents in soil profiles for several days of a drying period…………...
Estimated unsaturated hydraulic conductivity of soil matrix below 60 cm…..
Estimated daily evaporation for dry periods during the experiment………….
Probability distribution function of estimated daily evaporation……………..
Relationships between pressure head and water content at the 15-cm depth
between 140 and 190 day from the beginning of the experiment…………….
Drying water retention curve of a soil with macroporosity……………..…….
Cumulative water fluxes computed with constant evaporation of 0.86 mm
day.……………………………………………………………………………
Time series of the depth-averaged water contents simulated with the
calibrated HYDRUS-1D model ………………………………………...……
Soil water fluxes simulated with calibrated Richards medium model………..
Design of the model abstraction application in this work. The Richards
media was simulated with the HYDRUS-1D software; the water budget
model was MWBUS…………………………………………………………..
Time series of the depth-averaged water contents simulated with the
calibrated MWBUS model……………………………………………………
Soil water fluxes simulated with calibrated HYDRUS-1D and the MWBUS
models………………………………………….……………………………...
F-statistics of soil water flux simulations with different abstraction
techniques for depths of 15 and 55 cm………………………………………..
Time series of transect-average water contents: symbol – measured, red lines
– calibrated the HYDRUS-1D model, blue lines - calibrated the WMBUS
model for homogeneous soil profile…………………………………………..
Cumulative water fluxes simulated with calibrated HYDRUS-1D and
MWBUS models for homogeneous soil profile ……………………………...
Comparison of field water retention (symbols) at the depth of 15 cm with
water retention estimated using 21 pedotransfer functions listed in the
Appendix A and the Rosetta software………………………………………...
Time series of transect-average water contents: symbol – measured, red lines
- simulated with the HYDRUS-1D, blue lines - simulated with the WMBUS
model; water retention data and median hydraulic conductivity estimated
with pedotransfer functions…………………………………………………...
Soil water fluxes simulated with HYDRUS-1D and MWBUS models; water
retention data and median hydraulic conductivity estimated with
pedotransfer functions for layered soil profile………………………………..
Time series of transect-average water contents: symbol – measured, red lines
- simulated with the HYDRUS-1D, blue lines - simulated with the WMBUS
model for homogeneous soil profile; water retention data and median
hydraulic conductivity estimated with pedotransfer functions for topsoil……
Soil water fluxes simulated with HYDRUS-1D and the MWBUS model;
water retention data and median hydraulic conductivity estimated with
pedotransfer functions for topsoil profile; error bars show one standard
x
78
79
80
81
82
84
84
85
86
87
89
91
94
95
96
97
100
101
102
104
105
107
3-39
3-40
3-41
3-42
3-43
3-44
3-45
deviation………………………………………………………………………
Soil water fluxes simulated with calibrated HYDRUS-1D and the MWBUS
model for the homogeneous soil profile………………………………………
Schematics of the neural network application to abstract simulation models...
Estimated with the artificial neural network and simulated monthly soil
water fluxes at the depth of 105 cm. Model – MWBUS, artificial neural
network input – daily average flux, ANN output – daily average flux, the
total number of hidden neurons – 20………………………………………….
Estimated with the artificial neural network and simulated monthly soil
water fluxes at the depth of 105 cm. Model – MWBUS, artificial neural
network input – daily average flux, ANN output – daily average flux, the
total number of hidden neurons – 12……………………………………...
Estimated with the artificial neural network and simulated monthly soil
water fluxes at the depth of 105 cm. Model – Hydrus-1D, artificial neural
network input – daily average flux, ANN output – daily average flux, the
total number of hidden neurons – 20………………………………………….
Estimated with the neural network and simulated monthly soil water fluxes
at the depth of 105 cm. Model – HYDRUS-1D, ANN input – weekly
average flux, ANN output – weekly average flux, number of hidden neurons
– 12………………………………...………………………………………….
Probability distribution of the root-mean square error (RMSE) of water
content from Monte Carlo simulations for the HYDRUS-1D and the
MWBUS models with two sources of soil hydraulic properties and two
representations of soil profile…………………………………………………
108
109
111
112
113
114
115
117
Tables
2-1 A set of model simplification techniques suggested by Innis and Rextad
(1983) for four steps of model development.……………………..……………
2-2 Meaning and theoretical basis of model abstraction techniques……………….
2-3 Regression equation to predict soil water content at specific matrix potential
(after Rawls et al., 1982)……………………………………………………….
2-4 Regression equation to predict soil water content at specific matrix potential
(after Rawls et al., 1983)……………………………………………………….
2-5 Saturated hydraulic conductivity (cm day-1) at three probability levels for
USDA textural classes (from Rawls et al., 1998)……………………………...
2-6 Parameters of the Ahuja equation (2-3) to estimate saturated soil hydraulic
conductivity…………………………………………………………………….
2-7 Potential need for model abstraction from the parameter uncertainty
standpoint based on Hill (1998) analysis………………………………………
3-1 Average soil texture (%) along the trench at the Bekkevoort site……………..
3-2 Types and frequencies of the monitoring measurements………………………
3-3 Coefficients of variation (%) of soil water contents in time…………………...
3-4 Coefficients of variation (%) of soil water content along the trench…………..
3-5 Statistics of van Genuchten parameters of water retention in samples along
the 30-m rench…………………………………………………………………
3-6 Calibrated van Genuchten (Eq. 2-2) and van Genuchten-Mualem (Eq.2-5)
xi
6
8
27
28
29
32
41
53
60
65
65
70
parameters of the Richards medium model (HYDRUS-1D)………………….. 90
3-7 Calibrated parameters of the water budget model (MWBUS)………………… 98
3-8 PTF used to run HYDRUS 1D with Brooks-Corey (BC) and Van Genuchten
(VG) water retention equations………………………………………………... 103
xii
EXECUTIVE SUMMARY
This report describes and demonstrates the methodology of model abstraction in
subsurface contaminant hydrology. Model abstraction is defined as methodology for
reducing the complexity of a simulation model while maintaining the validity of the
simulation results with respect to the question that the simulation is being used to
address. ). Thus, model abstraction reduces the complexity of the system to be simulated
to its essential components and processes through a series of conceptualizations, selection
of significant processes and appropriate scales, and identification of the associated
parameters.
The need in model abstraction addresses the concerns of the NRC licensing staff that the
use of overly complex simulation models in decommissioning plans and performance
assessment may cause an excessive burden of data collection and computations as well as
difficulties in interpreting simulation results and conveying the simulation approach to
both technical and lay audiences. The presumed risk of leaving some important process
or feature out often leads the model users to employing fairly complex flow and transport
models that simulate subsurface systems through detailed numerical grids and associated
data inputs, thereby introducing large computational and intensive data-collection
requirements. However, many of the detailed features, events and processes represented
in these complex models may have limited influence on the performance of a specific
site.
The feasibility of model abstraction has been demonstrated in many research and
engineering fields that give ample examples of models having strikingly different
complexity and yet the same accuracy. In contaminant hydrology, model abstraction is
possible because the complexity of flow and transport pathways at a specific site is easy
to perceive but difficult to represent in mathematical equations of the model without
making strong simplifying assumptions. Different sets of plausible assumptions lead to
different models that are consistent with the available observations. The multiplicity of
models reflects the multiplicity of possible conceptual approaches to representation of
complex subsurface processes in mathematical form tractable within limitations of
existing computer and measurement technologies.
The methodology of model abstraction has been developing for more than 30 years in
various research and engineering fields. Most of the model abstraction techniques are
specific to the type of mathematical models used in a specific field. This report contains
an overview and annotated examples of model abstraction in modern contaminant
hydrology. Two main targets of abstraction are (1) the model structure, i.e. the formal
description of specific processes and their interactions that affect flow and transport
variables, and (2) the parameter determination, i.e. the estimation of constant and
functions serving as coefficients in model equations.
Structure of hydrologic models is changed with model abstraction via (a) using predefined hierarchies of models, (b) delimiting input domain, (c) scale change done by
either upscaling or aggregation, and (d) metamodeling. A predefined hierarchy contains a
xiii
series of progressively more simple conceptual and corresponding mathematical
representations of porous media. The class of model abstraction techniques based on the
delimiting input domain utilises the fact that some features, events, or processes may be
not relevant for a given set of scenarios or for a given set of model outputs. Scale change
provides transitions between four operational scales – core, profile/”pedon”, field, and
watershed scales that are of interest in contaminant hydrology. Model abstraction with
scale change alters model equations, variables and parameters with two classes of
methods: upscaling and aggregation. Upscaling model abstraction methods use the finescale model and the fine scale media properties to derive the coarse-scale model
equations and to relate the coarse-scale and fine-scale transport parameters. On the
contrary, model abstraction with aggregation does not assume any relationship between
model parameters and equations at the fine and at the coarse scales. Parameters of the
coarse-scale model are lumped, and are subject to calibration with field data. Aggregation
can be also done without the change in the model equations by combining several soil
horizons or geologic strata. Metamodeling seeks to simulate the input-output
relationships of the complex model with a simple statistical relationship.
Model abstraction applied to parameterization of hydrologic models does not provide a
substitution for the model calibration, but rather seeks to obtain reasonable estimates of
parameter values and their variability. Such estimates are useful for setting initial values
of parameters for the model calibration, using non-calibrated model in pilot studies and
field campaign designs, assigning values of parameters that are shown to be not sensitive,
etc. Abstraction of parameter determination affects parameter estimation via (a)
discretization, (b) scaling, and (c) pedotransfer functions. The discretization abstraction
techniques replace the continuously varying spatial fields of parameters with piece-wise
constant spatial distributions. The scaling abstraction techniques are useful when models
are coarsened, i.e. the grid cell size is substantially increased but no changes in the model
structure is made. The scale-dependence in model parameters is usually encountered in
such cases, and scaling relationships are needed to convert the original parameter values
and measurement data to the parameters values of the coarsened model. The pedotransfer
techniques convert the readily available data to the hydraulic and transport parameters of
the unsaturated flow and transport. Those parameters are notoriously difficult to measure,
and s substantial effort has been made to estimate these parameters from the data
available from soil survey or borehole logs. The empirical functions used for such
estimating are often called pedotransfer functions. Large databases have been assembled
to encompass variety of soil properties in developing pedotransfer functions in different
parts of the world. New powerful heuristic tools, such as artificial neural networks and
regression trees, appeared to be useful in the development of pedotransfer functions. Still,
the accuracy of pedotransfer functions outside of their development dataset remains
essentially unknown, and the use of ensembles of pedotransfer functions is strongly
recommended.
The model abstraction implementation has to conform requirements of objectiveness,
systematic implementation, comprehensiveness, and efficiency. The MA process starts
with an existing base model that can be calibrated and used in simulations. The key
output of the model id defined that provides the necessary and sufficient information to
xiv
decide on issues of interest. The base model may need abstraction for one or more of the
following reasons (a) the base model includes a complex description of processes that
cannot be observed well and yet need to be calibrated; the calibrated values of parameters
of those processes are very uncertain, (b) the base model propagates uncertainty in the
initial distributions, parameters, and forcing in a manner that creates an unacceptable
uncertainty of the key output, (c) the base model produces inexplicable results in terms of
the key output, (d) the base model requires an unacceptable amount of resources for
computations, data preprocessing, or data post-processing, e.g. the base model is not
suitable to be used as a part of a real-time modeling system that requires short computer
runtimes, (e) the base model lacks transparency to be explicable and believable to the
users of the key output.
The model abstraction process, as described in this report, is a transparent step-by-step
formalized procedure of justification of the use of a simplified model. It includes the
following steps (1) justify the need for the model abstraction, (2) review the context of
the modeling problem, (3) select applicable model techniques, (4) determine the model
abstraction directions that may give substantial gain, (5) simplify the base model in each
direction. Statistical criteria are provided to justify of the need in model abstraction in
case it is related to the uncertainty in calibrated parameter values or in the model key
output. The context of the modeling problem is described that has to be reviewed to
realize what details and features of the problem are omitted or de-emphasized when the
abstraction is performed, and thus to warrant the comprehensiveness and objectiveness of
the model abstraction process. Model abstraction can lead to simplifications via (a) the
number of processes being considered explicitly, (b) process descriptions, (c) coarsening
spatial and temporal support, (d) the number of measurements for the reliable parameter
estimation, (e) reduced computational burden, (f) data pre-processing and postprocessing. To guide the selection of applicable model abstraction techniques, classes of
model abstraction techniques are specified that lead to each type of the simplification.
Each abstracted model has to be parameterized and confirmed in the uncertainty context.
The model abstraction case study was design to illustrate the model abstraction process
and techniques. The main objective of the test case was to demonstrate how the model
abstraction can be applied to understanding and prediction of soil water fluxes at a
relatively humid site where transport may be affected by the presence of soil macropores
and related preferential flow phenomena. The test site was located in Bekkevoort,
Belgium. The 1.5-m deep test trench was dug in a loamy soil and was instrumented to
measure soil water content and soil matric potential in 60 locations at five depths hourly,
soil temperature in 6 locations at five depths hourly, and soil water fluxes in three
locations at two depths once in one to four days. Vegetation was removed and a thin layer
of fine gravel covered the soil. Soil water monitoring continued for 384 days. Soil was
sampled at 90 locations at three depths to measure soil water retention in laboratory. The
base model was the Richards equation of water flow in variably saturated porous media.
The model was calibrated using almost 200,000 measurements of soil water content and
soil water potential. The key output of the model was total soil water flux at two
measurement depths and at the bottom of the soil profile over three wetting-drying
periods.
xv
The need in model abstraction was justified by the fact that the calibrated base model
predicted substantial runoff whereas no runoff was observed at the site during the
monitoring period. The base model produced inexplicable results in terms of the key
output.
The review of the modeling context revealed that (a) soil water contents across the trench
demonstrated strong temporal persistence, and (b) a substantial preferential flow should
occur in soil. The persistence was used to estimate soil bottom flux from water content
measurements, and those flux flow estimates were used to assess evaporation from soil
surface. Thus, the review of modeling context helped to improve knowledge about
boundary fluxes that have not been measured.
Four classes of applicable model abstraction techniques were selected. Model structure
abstraction was achieved using the hierarchy of model, aggregation, and metamodeling.
The hierarchy of models included the Richards equation and a simple soil water budget
model. Aggregation was done by replacing the layered soil profile with the homogeneous
profile. Metamodeling was done with the artificial neural network that used daily and
weekly precipitation as input to estimate monthly cumulative soil water fluxes. An
ensemble of 22 pedotransfer functions was used in model abstraction of parameter
determination. HYDRUS1-D, MWBUS, and MATLAB software packages were used to
calibrate the abstracted models and to run Monte Carlo simulations to evaluate the
uncertainty in calibrated model outputs.
The abstraction with the hierarchy of models was useful. The simple soil water budget
model was less accurate in predictions of soil water content compared with the Richards
model. However, unlike the more complex Richards model, this simple soil water budget
model correctly predicted the absence of runoff and measured cumulative soil water
fluxes. The prediction of runoff was an artifact of the Richards model calibration in
absence of measured boundary fluxes. This abstracted model appeared to be instrumental
in both explaining behavior of the complex model and in predicting the key output – soil
water fluxes.
The abstraction with aggregation was not useful in this case. The Richards model was
less accurate with respect to soil water contents and continued to generate large simulated
runoff when a homogeneous soil layer was introduced.
The abstraction of parameter determination with pedotransfer functions was useful. It
showed that the Richards model with parameters in correct ranges is able to correctly
simulate soil water fluxes. The ensemble of pedotransfer functions represented field
water retention better than data from laboratory soil water retention measurements. Soil
water content predictions with the Richards model were more accurate with the ensemble
of pedotransfer functions than with laboratory data.
xvi
The neural network metamodel was extremely accurate in estimating cumulative soil
water fluxes. It was also five orders of magnitude faster than the numerical algorithm
coded in HYDRUS-1D.
In general, the model abstraction process was successful in this case study. Model
abstraction explained the strange behavior of the complex model, and provided the
correct description of the system behavior and plausible parameter ranges. The case has
also demonstrated the need in model calibration and uncertainty analysis toolboxes and
appropriate user interfaces to perform the model abstraction in the uncertainty context.
Overall, the intensive use of models in subsurface contaminant hydrology has resulted in
development of many model abstraction techniques that to-date have been used mostly in
research. Potential benefits of model abstraction include improvement of understanding
and communication of modeling results, more robust predictions, and better
understanding of essential factors and their representation in models. This makes model
abstraction an attractive methodology for engineering modeling applications. The model
abstraction process can be set as a transparent step-by-step formalized procedure of
justification of the use of a simplified model. An important feature of models abstraction
is the explicit treatment of model structure uncertainty. The model structure, along with
the data uncertainty, and scenario uncertainty, is known to introduce the uncertainty in
modeling results. Unlike the uncertainty in input data, in model parameters, and in
scenarios, the effect of the model structure uncertainty on the uncertainty in simulation
results is usually impossible to quantify in statistical terms. Using model abstraction, a
series of models with feasible structures can be built and evaluated in a systematic
manner. Each of the models is evaluated from results of an ensemble of simulations by its
accuracy to measurement data and by its predictions with respect to scenarios that have
not been observed.
It is expected that as the model abstraction methodology will develop, knowing about
model abstraction and applying it to the regulated sites will be helpful for the NRC
licensing staff. Reviewers will have a means to determine whether the models submitted
to support licensing actions adequately represent the site, and whether the investigations
adequately represent important features, events, and processes for the site. Managers will
have an advisory tool to decide whether the requests for additional information are
targeted tat sensitive site parameters and processes. On the other hand, the NRC licensees
will have a device to determine whether a simple model that is easy to understand and
communicate to regulators, stakeholders, and the general public can adequately represent
their site.
xvii
ACKNOWLEDGMENTS
This work was supported by the Interagency Agreement RES-02-008 “Model Abstraction
Techniques for Soil Water Flow and Transport.” The Nuclear Regulatory Commission
staff conceived and championed this research.
The authors acknowledge with thanks contributions by the following individuals: Dr.
Diederik Jacques of the SCK-CEN, Belgium for having provided experimental data from
the Bekkevoort experimental site; Dr. Brent Thomas of RAND Corporation for having
brought to our attention the recent extensive work on model abstraction; Professor
Shlomo Neuman of the University of Arizona and Dr. Glendon Gee from the Pacific
Northwest National Laboratory for their insightful review of a draft version of this
manuscript; members of the Working Group II on Uncertainty and Parameter Estimation
of the Interagency Steering Committee on Multimedia Environmental Models for their
helpful comments on the presentations of this work being in progress.
xix
Abbreviations
1D, 2D
ANN
ANOVA
AQUA
ASCII
CEC
CV
FAO
GA
HYDRUS
MA
MACRO
MEPAS
MIN3P
MODFLOW
MWBUS
Neuropack
NRC
PCAPS
PEST
PTF
PVC
Rosetta
SUFI
TDR
UCODE
USDA
WGEN
One-dimensional, Two-dimensional
Artificial neural networks
ANalysis Of VAriance between groups
One-dimensional water budget based model in unsaturated
soil
American Standard Code for Information Interchange.
Cation Exchange Capacity
Coefficient of variation
Food and Agriculture Organization of the United Nations
Genetic algorithms
Software package for simulating the movement of water,
heat and multiple solutes in variably saturated media.
Model Abstraction
A model of water flow and solute transport in Macroporous
soil.
Multimedia Environmental Pollutant Assessment System
Multicomponent reactive transport modeling in variably
saturated porous media
Three-dimensional finite-difference ground-water flow
model
Model of Water Budget of Unsaturated Soil
Neural Networks Package for fitting Pedotransfer Functions
U.S. Nuclear Regulatory Commission
Passive Capillary Samplers
A nonlinear Parameter Estimation and optimization package
Pedotransfer function
Polyvinyl Chloride
Windows-based program that implements artificial neural
network to soil hydraulic parameters
Parameter estimation methodology
Time Domain Reflectometer
A Computer Code For Universal Inverse Modeling
United States Department of Agriculture
A weather generator
Symbols
Roman symbols
a
A
b
B
C
C
CA
a parameter of PTF
a parameter of PTF
a parameter of PTF
a parameter of PTF
clay
coarse
clay activity
xx
cL
clay
D
ds
E
F
h
hb
he
hes
K
Ksat
Kw
l
L
lS
m
M
MF
n
N
OC
OM
Q
R
R2
RMSE
S
S
sand
SAR
sC
scL
Si
siC
sicL
siL
silt
sL
sβi
t
topsoil
VF
w
w330
w15000
clay loam
percent clay
mean depth of sample in centimeters
geometric mean diameter of soil particle size distribution
simulated daily evaporation rate
fine
capillary pressure (the absolute value of the matric potential)
bubbling pressure in the Brooks-Corey water retention model
air entry potential
air entry potential at a standard bulk density
number of the working probes at one depth
saturated hydraulic conductivity
unsaturated hydraulic conductivity
a parameter of unsaturated hydraulic conductivity model
loam
loamy sand
a parameter of van Genuchten’s water retention model
medium
medium fine
a parameter of van Genuchten’s water retention model
number of probes
percent organic carbon
percent organic matter
interlayer water flux
rainfall rate
Coefficient of determination
Root Mean Square Error
sand
soil water saturation
percent sand
sodium adsorption ratio
sandy clay
sandy clay loam
silt
silty clay
silty clay loam
silt loam
percent silt
sandy loam
standard derivation of the relative water content
time
a parameter of PTF
very fine
gravimetric water contents
gravimetric water contents at capillary pressures of 330 cm
gravimetric water contents at capillary pressures of 15000 cm
xxi
z
depth
Greek symbols
α
θj
a parameter of van Genuchten’s water retention model
relative water content
average relative water content
a parameter of unsaturated hydraulic conductivity model
water content change
time interval
thickness of the soil layer
a parameter of unsaturated hydraulic conductivity model
volumetric water content
effective porosity
water field capacity
average water content
θˆ j
average water content over working probes at this depth for the sampling time
j
θr
θs
λ
ρb
σg
τ
τ0
φ
residual water content
saturated water content
pore size distribution index in Brooks-Corey water retention model
soil bulk density
geometric standard deviation of soil particle size distribution
resistance parameter
the value of resistance parameter at depth z=0
porosity in Brooks-Corey water retention model
βij
β
γ
∆θ
∆t
∆z
δ
θ
θe
θFC
xxii
1 INTRODUCTION
1.1 Background
NRC staff review of performance assessments of nuclear facilities, e.g., decommissioning
of the facilities, management of low-level and high-level radioactive waste disposal sites,
frequently involves the review of models of water flow and solute transport in soils,
sediments, and porous rocks in the vicinity of the nuclear facility (U.S. NRC, 1993,
1994). These models seek to represent complex and highly transient subsurface systems.
Representation of those complex systems in existing models ranges from a very simple to
a extremely sophisticated (Davis et al., 2004; Meyer et al., 1997; Neuman and Wierenga,
2003; Reilly and Harbaugh, 2004; Hill et al., 2004).
It is generally agreed that models of the appropriate complexity may be required to
capture certain flow and transport phenomena (Hunt and Zheng, 1999). How to reach this
appropriate complexity remains an ongoing research issue. Suggested systematic
approaches are based on the scale of interest. A “downward” approach in hydrologic
modeling typically starts with a very simple model to capture the signature of, say, the
annual runoff response of a catchment (Milly, 1994). Complexity is added as the spatial
and temporal scales become finer. (Jothityangkoon et al., 2001). Neuman and Wierenga
(2003) addressed the issue of the suitable level of model complexity in the site-specific
hydrogeologic modeling. They indicated that site-specific hydrogeologic heterogeneities
and driving mechanisms may be controlled by the features, events and processes
belonging to the larger scale(s), and that the models have to account for these controls. At
the same time, site-specific hydrogeologic heterogeneities and driving mechanisms may
be also significantly influenced by the smaller scale features, events and processes, and
this also has to be accounted for in the model. Usually, the complexity of flow and
transport pathways at the specific site may be easily perceived, but it is often difficult to
represent it in mathematical equations of the model without making strong simplifying
assumptions (Beven, 2002). This implies that several different structures of the model
could be consistent with the available observations. Such multiplicity of models reflects
the multiplicity of possible conceptual approaches to representation of complex
subsurface processes in mathematical form tractable within limitations of existing
computer and measurement technologies.
It has been amply demonstrated that the increase in complexity of the model does not
necessarily mean the increase in its accuracy. Examples of equally accurate models of
strikingly different complexity can be found in various fields of modeling applications,
such as regional ground water assessments (Kelson et al., 2002), chemical engineering
(Diwekar, 1994), marine ecology (Stillman et al., 2000), runoff generation (Jakeman and
Hornberger, 1993), population dynamics (Stephens et al., 2002), demographic projections
(Smith, 1997), state-wide water supply systems (Palmer and Cohan, 1986), battlefield
simulations (Sisti and Farr, 1997), etc.
If the complexity is not inexorably linked with the accuracy, there may exist an
opportunity to simplify models. Simplifying models may stem from several rationales. As
1
motivations for model simplification, Neuman and Wierenga (2003) list narrowly defined
contextual or regulatory criteria, limited data or resources, and a quest for transparency.
The narrowly defined criteria usually mean that the accuracy of the model is evaluated
with respect to some (but not all) variables that this model predicts. Limitations of
resources may refer to runtime and also to data pre-processing and post-processing. The
transparency may be imperative in communicating model requirements and justifying the
use of the specific model.
1.2 Model abstraction
Model abstraction is the methodology of model simplification performed systematically
and objectively. Frantz (2002) defined the model abstraction as “a methodology for
reducing the complexity of a simulation model while maintaining the validity of the
simulation results with respect to the question that the simulation is being used to
address” (Frantz, 2002). Thus, the model abstraction reduces the complexity of the
natural system to be simulated to its essential components and processes through a series
of conceptualizations, selection of significant processes, and identification of the
associated parameters.
The presumed risk of leaving some important process or feature out often leads the model
users to employing fairly complex flow and transport models that simulate subsurface
systems through detailed numerical grids and associated data inputs, thereby introducing
large computational and intensive data-collection requirements. Many of the detailed
features, events and processes represented in these complex models may have limited
influence on the performance of a specific site. This will cause an excessive burden of
data collection and computations as well as difficulties in interpreting simulation results
and conveying the simulation approach to both technical and lay audiences. This has
become a concern on the part of both NRC staff and the staff of other regulatory and
research federal agencies involved in multimedia modeling. Applying model abstraction
in a systematic manner will address this concern.
Successful model abstraction results in a simpler model that renders more easily
understandable description of the problem and allows discussions to be focused on the
most important aspects. It is often imperative to explicitly acknowledge the abstraction
strategy to make a model justifiable and defensible. Finally, model abstraction explicitly
deals with uncertainties in modeling. Therefore, model abstraction techniques that can
simplify and expedite the assessment of complex systems without significant loss of
accuracy would greatly benefit the synthesis and review of performance assessments.
Recently, the issue of model simplification has been addressed for the subsurface
hydrologic modeling in the comprehensive strategy developed by Neuman and Wierenga
(2003) who advocated the need in systematic approach. This report develops and
illustrates such approach.
We note that the term “model abstraction” is sometimes applied to describe the model
development, i.e. representation of the complex reality with a specific conceptual and/or
2
mathematical model, or to describe a development of a set of progressively simplified
models. Such meaning of the term is not used in this report.
1.3 Objectives
The general goal of this work was to provide the NRC with an improved analysis
capability that would enhance the credibility and defensibility of site performance
reviews. Toward this goal, the specific objectives of this study were
(1) to identify model abstraction techniques that may be appropriate for characterizing
and simulating water flow and contaminant transport in vadose zone, and
(2) to develop a case study that will demonstrate the efficiency of some of model
abstraction techniques for a specific site.
Our presumption was that model abstraction techniques that reduce either the conceptual
complexity, or the parameter determination complexity, may also enhance
communications with stakeholders and help clarify the scientific bases for decisionmaking related to the first-order importance issues.
3
2 MODEL ABSTRACTION TECHNIQUES
2.1 Background
Two definitions of the term ‘simple’ are pertinent to modeling: (1) easily understood or
done, and (2) plain and uncomplicated in form, nature, or design (Weiner and Simpson,
1991). The definition 1 applies to the user perception and to the model use. The definition
2 pertains to the model structure and to the details of modeling. Simplification of models
has been analyzed in both directions for more than 30 years. The first comprehensive
discussion of model simplification was presented by Meisel and Collins (1973). These
authors listed four advantages of using the simplified models:
• they are less expensive to use; such savings will permit more thorough analysis
for a given analysis budget;
• they have fewer input requirements;
• they are easier to transfer and/or combine with other models;
• they are easier to interpret; it is easier to understand the properties of and results
from a model with a small number of state variables and parameters than the
properties and results of one with more.
The first summary of model simplification techniques was compiled by Ziegler (1976)
who introduced four categories of the model abstraction techniques:
• dropping unimportant parts of the model,
• replacing some part of the model by a random variable,
• coarsening the range of values taken by a variable,
• and grouping parts of the model together.
Innis and Rextad (1983) noted that model can be simplified at various stages of a
modeling project. They listed 17 categories of simplification techniques that were used in
ecological modeling (Table 2-1). The authors emphasized that the model simplification
has to be conditioned on the purpose of modeling and proposed the first guidelines of
using the model abstraction in modeling natural systems.
Frantz (1995) has viewed the model abstraction as a simplification of the model that
assures validity of the model for the specific purpose. He introduced a comprehensive
taxonomy of the model abstraction techniques presented in Fig. 2-1. Table 2-2 describes
the meaning of the classes of techniques proposed by Frantz (1995). The model boundary
abstraction changes the real-world factors accounted for in the model by changing the
exogenous variables of the model. The model behavior abstraction eliminates the
transitions through the system states that are irrelevant to the simulation requirements.
Finally, the model form abstraction uses inputs and outputs from the multiple simulations
with the model to define the relationship between input and output with a simpler
mathematical model. Frantz (1995) noted that different model abstraction techniques tend
to be applied simultaneously. Later, other classifications of model abstraction techniques
have been proposed (Caughlin and Sisti, 1997; Ziegler, 1998; Yilmaz and Ören, 2004)
that have less detail but generally cover the same set of techniques as the Frantz’ (1995)
classification.
5
Table 2-1. A set of model simplification techniques suggested by Innis and Rextad (1983)
for four steps of model development
No.
Technique
1
System
organization
2
Filtering
3
Stochastic
features
Graph theory
4
5
6
7
8
Sensitivitybased results
Structure and
logic analysis
Dimensional
analysis
Repro-Meta
modeling
9
Time
constants of
sub-systems
10
Analytical
solutions
Interdependen
cies
Perturbation
Calculation of
output
moments
11
12
13
14
15
Languages
Coding
improvement
16
Variance
reduction
Linear
systems
17
Examples
Step 1. Hypotheses
In a conservative system, one of compartment can be represented by
subtraction; self organization can impose relationships between parameters
that effectively eliminate one or more parameters,
Fluctuations in a state or driving variable at very high frequencies are not
sensible to the remainder of the system. For instance, rapid changes in solar
insolation have little effect on the growth of a tree.
Stochastic systems may be better represented directly rather than as
stochastically disturbed deterministic systems
Analysis of diagrams of the interrelations between variables may help to
isolate highly interactive subsystems for low resolution treatment
Step 2. Formulation.
Factors of little supposed importance may be omitted; susbsystems with little
effect on system dynamics can be aggregated.
Models often contain numerous model-building artifacts whose removal may
simplify the result.
Using dimensionless variables can make results more broadly applicable.
A model of a model may offer such advantages over the original model as (a)
finding interrelations not obvious in the data or the model, (b) simulated data
can be selected to meet restrictive requirements (e. g., being equally spaced in
time), (c) regression of input/output parts applied to all or part of simulation
may speed and reduce the cost of projection.
Widely disparate rates in a model may be eliminated using suitable
approximations, e.g. slowly changing variables may be replaced by constants,
rapidly changing variables that equilibrate with other variables may be
represented with algebraic expressions in place of dynamic equation, thus
eliminating a need in large number of parameters.
Using eigenvalues and eigenvectors instead of solving evolutionary matrix
equations
Interdependencies among parameters can be utilized that stem from data or
generally accepted theories.
Correct linearization of the initial model.
Computing statistical moments of the output directly from input data and
model structure
Step 3. Coding
Using the simulation language tailored to the needs of the study
Coding improvements can ease reading and debugging as well as increase the
execution speed.
Step 4. Experiments
Reducing the number of simulation runs needed to estimate an output value
within a specified accuracy; accounting for correlations between variables
Retaining the dominant eigenvalues and eigenvectors
6
Model abstraction techniques
Model boundary modification
Derive
Explicit
Hierarchy
of
models
Model behavior modification
Limited
Input
space
Approximation
Sensitivity
analysis
Probabilistic
input
Metamodeling
Causal
influences
Temporal
Unit
advance
Causal
decomposition
Look-up
table
and
interpolation
Selection
by
influence
State
Behavior
aggregation
Model form modification
Aggregation
of cycles
Event
advance
Function
Entity
By
function
By
structure
Numeric
representation
Figure 2-1. General taxonomy of model abstraction techniques after Frantz (1995). With
permission.
7
Table 2-2. Meaning and theoretical basis of model abstraction techniques (after Frantz, 2002)
Class of model
abstraction techniques
Hierarchy of models
Broad meaning.
Theoretical background
A pre-defined hierarchy of models can be created to include simplified
models and more complex models.
Limited input space
Limit the domain of inputs provided. Some process descriptions are not
important or can be simplified because input parameters are in a limited
range such that the additional processes have little or no effect on the
results.
Graph of models (Addanki et al., 1991);
compositional modeling (Falkenhauer
and Forbus, 1991).
Pre-defined experimental frame of
reference (Zeigler, 1976, Zeigler et al.
1998).
Approximation
Parameter elimination based on requirements of the simulation to be
executed rather than being defined externally. Analysis determines which
parameters in a model can be eliminated, yielding a new and
computationally simpler model that provides approximate results that are
within a certain tolerance necessary to meet the requirements of the
simulation
Causal relationships between parameters
(Nayak, 1992), aggregation of variables
(Simon and Ando, 1996).
Selection by influences
Simplifying models by using relationships between variables observed in
the simulations. This approach focuses on determining which variables in
the model must be exogenous, and requires that the effects of variables on
other variables in the model be defined.
Automated modeling (Rickel and Porter,
1994).
Behavior aggregation
Combining system states whose distinctions are irrelevant to the
simulation output. Temporal ordering of events may be irrelevant to the
final simulation result.
Automated modeling (Rickel and Porter,
1994).
Causal decomposition
Dividing a model into loosely connected components, executing each
component separately, and searching for constraints that execution of one
component can impose on other components. Where no causal relationship
exists, the components may be executed in parallel.
Tractable simulations (Clancy and
Kuipers, 1994), causality in model
abstraction (Iwasaki and Simon, 1994).
8
Aggregation of cycles
Combining states that reflect similar sequences and distinctions among the
individual sequences are irrelevant to the final outcome.
Aggregation in simulations (Weld,
1986).
Numerical
representation
Replacing continuous variables by categorical or nominal variables. For
example, combining states characterized by floating point numbers that
round to the same integer value.
Developments in genetic algorithms
(Mitchell, 1996) and in regression trees
(Clark and Pregibon, 1992).
Temporal aggregation
Decreasing temporal resolution. This type of abstraction reduces
computational complexity by reducing the granularity (coarseness) of the
time advance. This abstraction can be applied in either unit advance or
event advance in discrete event simulations
Temporal grain size aggregation
(Iwasaki, 1990)
Entity aggregation
Refers to the representation of a collection of “lower level” entities by a
“higher level” entity. Entity aggregation may be performed by structure or
by function. Aggregating units that perform common or similar functions
creates an entity that performs a function that is the aggregate of the
individual functions. Aggregation by structure is accomplished by
replacing lower level entities by higher level entities. Behavior in such
cases is substantially different as a result of abstraction, even if the output
is the same. Aggregates functions of entities or subunits but leaves entities
in place. Functions are redefined to provide a coarser list of states or
output information from existing entities.
Qualitative simulations (Fishwick and
Luker, 1991), variable resolution
modeling (Hillestad et al, 1992, 1993).
Look-up tables
Simplification of the input-output transformation within a model or model
component by means of a decrease in computational effort.
Probabilistic input
Using randomly generated inputs to develop lumped models.
Fishwick, 1988.
Metamodeling
Approximating the output response surface of the model with a statistical
regression or using heuristic data mining tools.
Neural network-based metamodeling
Burton (1994)
9
Recent developments in the model abstraction theory and applications include using the
multiresolution modeling as proposed by Fishwick (1995). This type of modeling
presumes existence and use of two or more models of different complexity for the same
system. Davis and Bigelow (2003) indicated that models of different resolution, or detail,
have different strengths and weaknesses. Detailed models are particularly valuable for
representing explicitly the underlying phenomena. Successful models of this type are the
embodiments in mathematics and/or in computer programs of the fundamental
knowledge about the subject in question existing at the moment of the model
development. However, the fundamental knowledge about the system under
considerations not the only important knowledge (or even the most important knowledge)
for the decision-making A significant part of our knowledge of the world is lowresolution. Bigelow and Davis (2003) noted that both the insightful strategy-level
analysis and the decision support typically require relatively simple models. The most
fundamental reason is cognitive: decision makers need to reason about their issues and
inject their own judgments and perspectives. They need to construct coherent stories that
are convincing to themselves and to others. This implies abstracting, that is reducing of
what may be a very complex problem to a relatively small number of variables or
cognitive chunks (e.g., 3–5 rather than 10s or 100s). This also implies focusing on the
appropriate variables and on the cause-effect relationships. Another fundamental reason
for using simpler models is that strategy-level problems may be characterized by massive
uncertainty in many dimensions. The appropriate way to address such problems is often
the exploratory analysis in which one examines issues across the entire domain of
plausible initial states. The exploratory analysis, however, is most effective when the
model comprehensively covers the space in which the problem is considered with only 3–
12 variables. In such cases, the exploration can be comprehensive and comprehensible. In
contrast, if one has a large model and explores by holding hundreds of variables constant
while varying only a few (because in reality it is not possible to vary all of them), the
results cannot be assessed confidently because the effects of varying the others are
unknown. Model abstraction also seems to be very attractive because simple models
require much smaller amounts of the experimental data, much less time and efforts for
data preparation and post-processing. Analysts and end users can quickly comprehend
simple models and their inputs and outputs.
Model simplification has also been recently advocated as a powerful tool to improve the
usefulness of complex ecological models for uncovering mechanisms of ecosystem
functioning (Van Ness and Scheffer, 2005). Thes authors have proposed a strategy that
includes joint exploration of the behavior of a complex model and of its simplified
counterpart. The purpose of the simplification in this case is to gain understanding about
causes of the patters in the complex model output.
Types of models applied in various research and engineering fields are quite different.
Therefore, not all model abstraction techniques are relevant to the particular field of
knowledge. For example, abstracting models of the discrete events (i.e. the event
advance abstraction in Frantz’s classification) is not relevant to subsurface hydrology in
its current state where the representation of flow and transport is predominantly
continuous. The arsenal of model abstraction techniques developed in linguistics or in
10
computer science has a very limited use in subsurface hydrology because of fundamental
differences in types of models used.
The need for the model simplification in contaminant hydrology was outlined by Neuman
and Wierenga (2003) in their comprehensive strategy of hydrologic modeling and
uncertainty for nuclear facilities and sites. Neuman and Wierenga emphasized the
necessity to carry out the model simplification a systematic and comprehensive manner.
In this work, simplifications of model structure and of model parameter determination are
considered. Simplifications related to the numerical solution of model equations are
outside the scope of this report.
2.2 Model abstraction techniques for flow and transport models
This section present a systematic description of the model abstraction techniques applied
in subsurface hydrology. The categories of model abstraction techniques relevant to flow
and transport modeling are presented in Fig. 2-2. This classification is based on our
personal experience and the extensive literature review that is given below. Although an
effort has been made to compile a comprehensive list, the classes of techniques outlined
in the next section are neither mutually exclusive nor mutually exhaustive. Two large
targets of abstraction are (1) the model structure, i.e. the formal description of specific
processes and their interactions that affect flow and transport variables, and (2) the
parameter determination, i.e. the estimation of constant and functions serving as
coefficients in model equations. Although model calibration is considered to be a
mandatory procedure in flow and transport modeling applications, a preliminary
estimation of model parameters and their variability parameters is useful in both setting
initial values of parameters for the model calibration and using non-calibrated model in
pilot studies and field campaign designs. Each category of the classification in Fig. 2-2
represents a variety of specific procedures and techniques that are identified in
following sections.
2.2.1 Hierarchies of models
A pre-defined hierarchy of models has been suggested for flow and transport in structured
media (Altman et al., 1996). Figure 2-3 shows a schematic representation of increasingly
complex models that may be used to simulate preferential flow and transport in
macroporous soils or unsaturated fractured rock. The simplest bucket-type model
describes the accumulation of water in soil matrix and its discharge when the water
content exceeds the field water capacity or during evaporation periods (Fig. 2-3a). The
classical approach for simulating flow/ transport processes in vadose zones devoid of
macropores or fractures is to use the Richards equations for variably-saturated water flow
and the advection-dispersion equation for solute transport (Fig. 2-3b). The simplest
situation (Fig. 2-3c) for a fractured medium arises when the Richards and advectiondispersion equations are still used in an equivalent continuum approach, but now with
composite (double-hump type) hydraulic conductivity (permeability) curves of the type
shown in Mohanty et al. (1997), rather than the classical curve for the relative
permeability shown in Fig. 2-3b. More involved dual-porosity type models (Fig. 2-3d)
11
Model abstraction
Abstraction of the model structure
Hierarchy
of models
Limited
input
domain
Scale
change
Upscaling
Abstraction of parameter determination
Metamodeling
Discretization
Pedotransfer
Scaling
Aggregation
Figure 2-2. Categories of model abstraction techniques relevant to flow and transport
modeling in subsurface hydrology.
12
a
c
d
e
Equivalent
Single
matrix and Dual porosity
contifracture
nuum
continuum Matrix Fracture
Dual
permeability
Matrix Fracture
f
Discrete
fractures
without
matrix
g
Discrete
fractures
with
matrix
Relative
permeability
Water
budget
b
Pressure head
Complexity
Model abstraction
Figure 2-3. Hierarchy of models to simulate water flow and solute transport in structured
soils or in unsaturated fractured rock. Modified from (Altman et al., 1996).
13
result when the medium is partitioned into fracture and matrix pore regions, with water
and/or solutes allowed to exchange between the two liquid regions (e.g., Ventrella et al.,
2000). Different formulations of this type are possible. For example, one could permit
transient variably saturated flow in the fractures only, while allowing water to exchange
between the fracture and matrix domains. The latter situation leads to both advective and
diffusive exchange of solutes between the fractures and the matrix but still without
vertical flow in the matrix (e.g., Zurmühl and Durner, 1996; Zurmühl, 1998). Dualpermeability models (Fig. 2-3e) arise when water flow occurs in both the fracture and the
matrix domains. Examples are models by Pruess (1991), Gerke and van Genuchten
(1993) and Jarvis (1999). These models use different formulations for the exchange of
water between the fracture and matrix regions. In some models (e.g., Wilson et al., 1998)
more than two domains are considered, each one having its own hydraulic properties.
The modeling approach can be refined further by considering transient flow and/or
transport in discrete fractures without (Fig. 2-3f) or with (Fig. 2-3g) interactions between
the fractures and matrix. The latter approach is based on the assumption that the flow and
transport equations of the fracture network can be solved in a fully coupled fashion with
the corresponding equations for the matrix (e.g., Therrien and Sudicky, 1996).
Subsets of this hierarchy that have been used most often include using two region and
one-region models of pore space (Nielsen et al., 1986). Other examples of hierarchies use
non-equilibrium and equilibrium models for contaminant sorption and transport,as
illustrated by Valocchi (1985).
2.2.2 Delimited input domain
This class of model abstraction techniques relies on the observation that some features or
processes may be not relevant for a given class of scenarios or for a given set of model
outputs. A reduction of the spatial dimensionality is one application of this technique.
Two-dimensional representations can be redundant and 1D representation may suffice as
shown by Wang et al. (2003). In another example, Guswa and Freuberg (2002) explored a
possibility to use the 1D model to characterize solute spreading in the environment with
low permeability lenses; they found that 1D macroscopic advective-dispersive equation
well matched the results from the 2D model when the equivalent conductivity of a
domain was less than the geometric mean conductivity, This example shows that one
should expect a change in parameter values if the dimensionality is reduced.
Delimiting input domain may allow rejection of the models in the hierarchy of Figure 2-2
based upon the type of flow being encountered. This is schematically illustrated in Fig.
2-4 (Nimmo, 2002) for the maximum rate of preferential flow over various distances for a
large number of field tracer experiments. The absence of saturated or near-saturated flow
may make it possible to use the equivalent continuum models, whereas the presence of
saturated flow may warrant the use of the dual-porosity or dual-permeability flow and
transport models. The presence of substantial periods of unsaturated flow and isolated
events of saturated conditions may motivate the use of different models for each of the
two types of flow. This use of two different models may be simpler than invoking one
integrated model for continuous simulation of both types of flow. Similarly, some types
14
Scale of transport (m)
Figure 2-4. Maximum tracer velocity in 34 documented field experiments with variably
saturated soils (after Nimmo, 2002, with permission).
15
of soil water regimes make the steady state flow assumption applicable to the solute
transport simulations at filed scale (Destouni, 1991).
The behavioral pattern of the biotic component may also allow abstraction of this type.
Guswa et al. (2002) have studied the soil water regime in presence of plants and have
found that the mechanistic soil water flow model can be abstracted to the simple bucketfilling model if the plant can extract water from locally wet regions to make up for roots
in dry portions of soil column.
2.2.3 Scale change
Scale is a complex concept having multiple connotations. A notion of support is
important to characterize and relate different scales in subsurface hydrology. Support is
the length, area, or volume for which a single value of porous media property is defined
and no variations in this and other properties are taken into account. Size of an individual
sample and size of a discrete spatial element in the flow and transport model present
typical examples of the support. The term "resolution" is used for supports defined in
terms of length, and the term "representative elementary volume" is applied for supports
defined as volumes. Terms "pixel size" and "grid size" are also used to define resolution.
An area or a volume that are sampled with given support determine the extent of
measurements. Distances between sampling locations define spacing. Blöschl and
Sivapalan (1995) suggested using the triplet support-spacing-extent to characterize the
scale of the hydrologic study. In practice of the vadose zone research, supports of core
scale, soil profile, or “pedon,” scale, field scale, and watershed scale are defined
operationally, and the increase of the linear size by about two orders of magnitude
corresponds to the transition from one of these scales to the next, coarser one (Fig. 2-5).
Change in spatial scale is usually reflected in the change of temporal scale as shown in
this figure.
The transition from a fine to the next, coarser scale in this sequence is referred below as
the scale change.
2.2.3.1 Scale change – upscaling
This category of model abstraction recognizes the need in altering model structure when
the spatial scale changes. Model equations, variables and parameters change.
Specific techniques of such abstraction employ statistical properties of flow and transport
at the fine scale to derive the coarse-scale models. Development of upscaling techniques
is very actively researched in hydrology, and a multitude of techniques is developed
(Govindaraju, 2002; Rubin, 2003). Most of the upscaling techniques use the some
assumptions about the spatial variability of flow, transport and porous media properties.
However, no standard technique exists to date. The differences among the techniques are
also related to the need of introducing additional, so-called constitutive relationships to
compensate for the loss of information that occurs after upscaling (Cushman, 1990).
16
1 cm
1m
core
0min
profile
30min 60min 0h
Steady-state
100 m
Event
12h
10 km
field
24h Jan
Jun
Seasonal
catchment
Dec
1980 1990
Decades
Figure 2-5. Spatial and temporal operational scales in hydrology (after Blöschl and
Sivapalan, 1995). With permission, John Wiley & Sons, Inc.
17
2000
Different constitutive relationships lead to different techniques. Recently it was suggested
to carry out upscaling in the probability space (Guadagnini and Neuman, 1999). With this
method, the effective parameters of the flow model depend on the available information
about the porous media properties at the fine scale.
Numerical stochastic simulations are also employed in upscaling. In this case the coarsescale flow and transport is estimated not from upscaled equations but directly from Monte
Carlo simulations at fine scale. Recently, using Monte-Carlo simulations became
instrumental in upscaling multiphase flow and transport models for unsaturated zone
simulations (i.e., Ewing et al., 2000; Li et al., 2003). Khaleel et al. (2002) used Monte
Carlo simulations of flow and transport to represent a heterogeneous unsaturated medium
with its homogeneous equivalent. Simulations showed that such upscaling also requires
introducing new constitutive relationships, such as dependences of hydraulic conductivity
anisotropy and dispersivity on moisture content or tension.
Upscaling may involve also the reduction of dimensionality, as demonstrated by Kumar
(2004) for upscaling two-dimensional Richards equation to its one-dimensional coarsescale analog.
To be efficient, all mentioned upscaling techniques have to use the correct statistical
representation of rare features at the fine scale, because these features often govern the
media behavior at the coarse scale. For example, if hydraulic conductivity is distributed
lognormally, then five percent of fine scale representative elementary volumes (REVs)
will conduct 95% of all flow, and relatively rare REVs will control the flow at the coarse
scale. Macropores are rare in soils and are easy to miss during sampling, but they control
water flow and solute transport in soil profile.
The scale finer than the core/core scale is usually referred to as the pore scale. This scale
is not of an immediate interest in engineering applications of the flow and transport
modeling Nevertheless, an active upscaling research at this scale is being done towards
better understanding and modeling of flow and transport at core scale with the possibility
to extend upscaling techniques to coarser scales. One example is to combine individual
stream tubes into a stochastic transport model (Jury, 1990). Recent developments include
modeling solute transport under assumption that individual solute particles undergo nonBrownian motion in a complex pore space. In this case, the classical convectivedispersive equation becomes inapplicable, and other models need to be used. Solute
particles in complex and/or reactive pore space may experience long waiting times before
making a “leap” in the general direction of transport. Such individual behavior can be
integrated using models of fractional kinetics (Benson et al., 2001) or more general
continuous-time random walk models (Berkowitz et al, 2001). A similar approach to
modeling of the flow in unsaturated soil is also plausible (Pachepsky et al, 2003).
Although these methods have been mostly researched to understand transport at the core
scale, they are increasingly applied at the coarser scales. For example, the diffusionlimited aggregation model provided a good description of solute transport in soils with
18
the distance-dependent dispersivity (Flury and Flühler, 1995). Fractional kinetics models
have been successfully used to simulate the MADE tracer experiment (Benson et al.,
2001).
2.2.3.2 Scale change – aggregation
Change of scale also leads to a change of the model governing equations and parameters
with these techniques. However, unlike in upscaling, no relationship is assumed between
model parameters at the fine and at the coarse scales. Parameters of the coarse-scale
model are deemed to be lumped, and field data are needed to calibrate such coarse scale
models. An example is using a bucket-type soil water flow model (Fig. 2-3a) at the field
scale whereas Richards equation is used at the soil profile scale (Fig. 2-3b). Soil water
retention is parameterized with the field capacity at coarse scale and with soil water
retention curves at the fine scale. No relationship between the coarse scale and fine scale
parameters exists, because there is no reliable relationship between the water content at
field capacity and soil water content at any specific values of soil water tension (Fig. 2-6).
Aggregation can be also done without the change in the model equations by combining
several soil horizons or geologic strata. In flow and transport in vadose zone, one
common application is to replace a heterogeneous soil profile with an equivalent
homogenous profile while retaining the Richards equation as a flow model (Zhu and
Mohanty, 2002). In such case, flow and transport parameters are still lumped. Attempts to
determine the effective hydraulic properties of the equivalent soil profile from the layer
properties have shown that the effective properties depend not only on layer properties
but also on the type of the predominant water regime (infiltration or evaporation). Many
examples of distributed watershed modeling with various degree of sub-watershed
aggregation were recently developed in surface hydrology. Boyle et al. (2001) showed
that aggregation from eight to three sub-watersheds does not worsen the model
performance whereas aggregating all sub-watersheds in one does.
Aggregation should be distinguished from the model coarsening, when the support is
increased without explicit recognition of the increasing non-homogeneity within it. In the
latter case, no model abstraction is actually performed.
2.2.4 Metamodeling
This is a group of abstraction methods that use results of multiple simulation runs to
extract the information helpful to simplify a complex model.
Metamodeling literally means modeling a model. Also known as the repromodeling
(Meisel and Collins, 1973) or the response surface modeling, the metamodeling creates a
relatively small and simple empirical model intended to mimic the behavior of a large
complex model, that is, to reproduce the object model’s input-output relationships. (Davis
and Bigelow, 2003). A common way to develop a metamodel is to generate “data” from a
19
(Water content at a fixed potential)/(field capacity)
2.0
1.5
CS
MS
FS
LS
CS LFS
MS
LS
FS
LFS
1.0
0.5
0.0
0
10
20
30
40
Water content at a fixed potential
Figure 2-6. Soil field capacity and soil water retention at fixed matric potentials; CS coarse sand, MS - medium sand, FS - fine sand, LFS - loamy fine sand at -5 and -10 kPa
sol matric potentials (data from Rivers and Shipp, 1978); ○ and □ - topsoil and subsoil of
varying texture at -33kPa (data from Haise et al., 1955).
20
number of large-model runs and then to use the statistical methods to relate the model
input to the model output without attempting to understand the model’s internal working.
Artificial neural networks (ANN) became a popular means to build metamodels because
they are powerful approximators and, as such, are used to relate multiple input variables
to outputs from the complex model (Hecht-Nielsen, 1990). Recent examples of
application of the artificial neural networks to mimic MODFLOW output in a range of
scenarios for particular remediation sites have been recently published (Kron and
Rosberg, 1998; Masket et al., 2000).
The use of ANN requires development of a large number of training datasets covering the
range of possible scenarios of forcing variables. Generating such datasets requires an
extensive computing effort, but the computing with a trained ANN is several orders of
magnitude faster than the simulation with the original model. Wang and Jamieson (2002)
reported an example of regional wastewater planning in which, for reasons of
computational efficiency, an artificial neural network was employed to replicate the
process-based model in multiple evaluations of the model output during optimization
aiming to determine both the best sites and individual discharge standards.
ANN can be used to simulate not only the model output of interest, but also the results of
any part of computations performed during a model run. Hassan and Hamed (2001)
demonstrated the use of ANN to predict particle trajectories in a particle-tracking
algorithm in simulation the plume migration in heterogeneous media.
Using classification and regression trees (sometimes called decision trees) is another
heuristic methodology for creating metamodels. It gains popularity because of the
transparency of results, ability to retain only significant inputs-predictors, and possibility
to work with categorical, or ordinal inputs. No applications of regression trees in
subsurface flow and transport modeling have been found, although they were
successfully used in hydraulic parameter estimation from readily available data (Rawls
and Pachepsky, 2002).
Based on pure statistics, metamodels undoubtedly have problems that may be either
minor or major, depending on the context and on the statistical methods used. Davis and
Bigelow (2003) listed (a) the failure to tell a story, that is inability of metamodels to
provide a sensible explanation of results to a decision-maker, (b) the failure to represent
multiple critical components; if a system failure is modeled and the system fails if any of
several components fail, then metamodels may not simulate that, instead predicting that a
weakness in one component can be compensated by greater strength of another, (c) nonreliable sensitivity of metamodels, and (d) shortcomings in representing the whole input
space; statistical metamodel may emulate base-model results reasonably well on average,
but fails badly in what appears to be the obscure corners of the input space. Those authors
suggest to use statistical metamodels based on the physical phenomena represented in the
object model, knowledge of the structure of the object model, or a combination.
21
2.2.5 Discretization
In some cases, using a set of discrete values of parameters instead of their continuous
values leads to substantial simplifications. This simple but very important abstraction is
commonly used in flow and transport modeling by employing soil and geological textural
maps to define soil or sediment "units" that have the same hydraulic properties - even
though the real-world hydraulic properties may gradually change from one unit to
another. Zonation of hydraulic conductivity was found efficient when geophysical and
hydrogeological data were combined (Hubbard et al., 2003). Automated zonation is now
feasible (Tsai, 2003) that gives an advantage of no need in assuming a specific equation
to simulate the dependence of parameters on the spatial coordinates.
2.2.6 Scaling
When the model coarsening is performed, i.e. the support is increased but no changes in
the model structure is made, the scale-dependence in model parameters will be most
probably encountered because the larger units of the hierarchical structures in soils or
sediments are enclosed in the extended support volumes. This kind of scale-dependence
has been documented for the groundwater flow (Neuman and Di Frederico, 2003). A
similar scale-dependence can be clearly observed in variably saturated soils as shown in
Fig. 2-7, 2-8, and 2-9. Whereas in groundwater studies scaling of dispersivity was shown
to obey the universal scaling law (Neuman, 1990, 1995), the scaling exponents of
dispersivity in unsaturated soils appear to be quite variable (Fig. 2-7). The scale-related
increase in saturated hydraulic conductivity depends on whether an aggregation of soil
horizons has actually occurred in experiments (Fig. 2-8). Substantial differences in coremeasured and field-measured water retention (Fig. 2-9) show that the applications of the
core scale unsaturated flow parameters at field scale require cautious consideration.
Scaling relationships for the parameters of the unsaturated flow have been documented
with the support range of about 1 to 1.5 orders of magnitude (Fig. 2-7, 2-8, and 2-9).
They can provide useful relationships for modeling within the core-pedon range of scales
if a model is coarsened, i.e. discretization in computations is changed without change in
model equations. In such case, scaling provides means of using core scale measurements
to infer parameters for increased support size
2.2.7 Parameter estimation with pedotransfer functions
Parameters of the water flow models in variably saturated porous media are nonlinear
functions of matric potential or water content. These parameters are notoriously difficult
to measure. Measurement of water and solute fluxes in unsaturated soils remains a
research issue; no routine methods have been implemented to-date. Therefore, defining
parameters using the inverse methods is often problematic. A substantial effort has been
made to estimate these parameters from the data available from soil survey or borehole
logs. The empirical functions used for such estimating are often called pedotransfer
functions. An extensive work has been done to develop them. Proceedings of two
International conferences (van Genuchten et al., 1992, van Genuchten et al, 1997) and the
22
Dispersivity, cm
10
1
0.1
10
100
1000
Distance, cm
Figure 2-7. Observed dependencies of the solute dispersivity on distance in unsaturated
soils (after Pachepsky et al., 2000). With permission, SSSA. Different symbols show data
of different authors.
23
-1
Saturated hydraulic conductivity (mm s )
10
1
0.1
0.01
102
103
104
105
Sample cross-section area (cm2)
Figure 2-8. Examples of scale-dependence in soil saturated hydraulic conductivity; ○loamy sand, Ap horizon, □ - loamy sand, A2 horizon, ∆ - sandy clay loam, B1t horizon, L
- clay loam, B1t horizon (after Pachepsky and Rawls, 2003). With permission, Blackwell
Publishing.
24
Field water content (cm3 cm-3)
0.6
0.5
0.4
0.3
0.2
0.1
0.0
0.0
0.1
0.2
0.3
0.4
0.5 0.6
Laboratory water content (cm3 cm-3)
Figure 2-9. Relationship between field and laboratory water contents at the same soilwater matric potential; { - samples with sand content more than 50% , z - samples with
the sand content less than 50%; lines show the quadratic regression ( ______ ) with 95%
prediction interval (_ _ _). Data from the UNSODA database (Leij et al., 1996). With
permission, SSSA.
25
first book on this topic (Pachepsky and Rawls, 2004) provide a panorama of these fastdeveloping studies.
2.2.7.1 Estimating water flow parameters
For water retention, one approach consists in estimating soil water contents at several soil
water potentials with a separate regression equation for each potential. Rawls et al. (1982)
used the US Cooperative Soil Survey Database to develop 12 regression equations to
relate soil water contents at 12 matric potentials to sand, clay and organic matter contents
(Table 2-3). Later, a similar set of equations has been developed to utilize the values of
bulk density along with sand, clay and organic matter contents (Table 2-4). Other
predictive equations developed from large databases are presented in the Appendix A.
Another approach to estimate soil water retention involves estimating coefficients in
equations that express dependence of soil water contents on soil water potential. The
equations that are most often used are the Brooks-Corey equation (Brooks and Corey,
1964):
λ
θ − θ r ⎛ hb ⎞
(2-1)
=⎜ ⎟
φ −θr ⎝ h ⎠
and the van Genuchten equation (van Genuchten, 1980):
θ −θr
1
=
(2-2)
θ s − θ r 1 + (αh) n m
[
]
In Eq. (2-1) and (2-2), θ is the volumetric water content, h is the capillary pressure (the
absolute value of the matric potential), φ is the porosity, θs is the saturated water content,
θr is the residual water content, hb is bubbling pressure, λ is pore size distribution index,
α, m, and n are empirical shape-defining parameters. Various equations to estimate
parameters in Eq. (2-1) and (2-2) are presented in the Appendix A. Recently, several new
water retention equations were introduced to describe θ-on-ψ dependencies (Kosugi,
1994; Fayer and Simmons, 1995; Perfect et al., 1996, Bird and Perrier, 2000) but they are
lacking the pedotransfer functions.
For the saturated hydraulic conductivity, Ksat , one approach is to estimate it from soil
basic parameters without employing data on soil water retention (i.e., Bloemen, 1980;
Loague, 1992). Rawls et al. (1998) assembled a database of more than 1000 experiments
in the USA and showed that both median value of Ksat and its values on 25% and 75%
probability levels can be defined for each USDA textural class if the samples are
preliminary separated into high porosity and low porosity groups for each of the classes.
These results are shown in Table 2-5. Referring the median Ksat values and the
difference between Ksat values at 25% and at 75 % probability levels to the average clay
content of the class results in a fairly well defined relationships (Fig. 2-10).
Another approach to the Ksat estimation benefits from soil pore space models as a source
of the Ksat equation in which coefficients need to be estimated from soil basic properties.
Ahuja and co-workers (1984) introduced a generalized form of the Kozeni-Karman
26
Table 2-3. Regression equation to predict soil water content at specific matrix potential:
θ(ψ) (cm3cm-3)= a + b*sand(%) + c*silt(%) + d*clay(%) + e*OM(%) (after Rawls et al.,
1982)
ψ, kPa
-10
-20
-33
-60
-100
-200
-400
-700
-1000
-1500
Parameters of the linear regression
a
b
c
d
0.4118 0-0.0030
0
0.0023
0.3121
-0.0024
0
0.0032
0.2576
-0.0020
0
0.0036
0.2065
-0.0016
0
0.0040
0.0349
0
0.0014 0.0055
0.0281
0
0.0011 0.0054
0.0238
0
0.0008 0.0052
0.0216
0
0.0006 0.0050
0.0205
0
0.0005 0.0049
0.0260
0
0
0.0050
27
e
0.0317
0.0314
0.0299
0.0275
0.0251
0.0200
0.0190
0.0167
0.0154
0.0158
Table 2-4. Regression equation to predict soil water content at specific matrix
pressure:θ(ψ) (cm3cm-3)= a + b*sand(%) + c*clay(%) + d*OM(%) + e*ρb(g cm-3) (after
Rawls et al., 1983)
ψ, kPa
-20
-33
-60
-100
-200
-400
-700
-1000
-1500
Parameters of the linear regression
a
b
c
d
0.4180 -0.0021 0.0035 0.0232
0.3486 -0.0018 0.0039 0.0228
0.2819 -0.0014 0.0042 0.0216
0.2352 -0.0012 0.0043 0.0202
0.1837 -0.0009 0.0044 0.0181
0.1426 -0.0007 0.0045 0.0160
0.1155 -0.0005 0.0045 0.0143
0.1005 -0.0004 0.0045 0.0133
0.0854 -0.0004 0.0044 0.0122
28
e
-0.0859
-0.0738
-0.0612
-0.0517
-0.0407
-0.0315
-0.0253
-0.0218
-0.0182
Table 2-5. Saturated hydraulic conductivity (cm day-1) at three probability levels for
USDA textural classes (from Rawls et al., 1998)
USDA textural
class
Sand
Fine sand
Loamy sand
Loamy fine sand
Sandy loam
Fine sandy loam
Loam
Silt loam
Sandy clay loam
Clay loam
Silty clay loam
Clay
Soils with high porosity
Porosity
25% 50% 75% cm3cm-3
638.4 436.8 230.4 0.44
566.4 338.4 283.2 0.49
468.0 295.2 201.6 0.45
292.8 148.8 86.4
0.46
312.0 134.4 72.0
0.47
86.4 52.8 24.0
0.45
67.2 9.4 3.8
0.47
89.0 34.6 18.2
0.49
121.2 18.5 4.8
0.44
31.4 10.1 5.3
0.48
25.0 8.9 5.5
0.50
14.4 4.8 2.2
0.48
29
Soils with low porosity
Porosity
25% 50% 75% cm3cm-3
523.2 218.4 153.6 0.39
528.0 240.0 163.2 0.39
184.8 98.4 74.4
0.37
278.4 28.8 16.8
0.37
74.4 31.2 0.37
0.37
40.8 19.7 8.2
0.36
39.6 14.9 6.7
0.39
23.8 8.2 2.4
0.39
26.2 6.7 2.4
0.37
9.1 1.7 0.5
0.40
33.6 11.8 5.5
0.43
16.6 4.3 0.7
0.40
Ksat(75%)-Ksat(25%) (cm d-1)
Ksat(50%) (cm d-1)
1000
100
10
1
1
10
100
1000
100
10
1
1
10
100
Clay content (%)
Figure 2-10. Saturated hydraulic conductivity Ksat grouped by textural classes and bulk
density classes; (a) median values Ksat(50%) for the groups, (b) difference between third
Ksat(75%) and first Ksat(25%) quartiles of the distributions within each group; data for high
and low bulk density samples are shown by filled and hollow symbols, respectively;
, -sand, , - fine sand, , - loamy sand, , - loamy fine sand, , - fine sandy loam,
,
- loam,
, - silt loam, ,
- sandy
sandy loam, ,
- clay.
clay loam, , - clay loam, , -silty clay loam, ,
30
model to relate the saturated hydraulic conductivity to the effective porosity:
n
K sat = Bθ e
(2-3)
where θe is a difference between the total porosity and soil water content at -33 kPa, B
and n are parameters which vary among different data sets (Table 2-6).
Estimates of the unsaturated hydraulic conductivity are predominantly based on pore
space models that relate unsaturated hydraulic conductivity to the saturated hydraulic
conductivity and to geometric parameters of the pore space inferred from water retention
data.
A general form of such model has been proposed by Alexander and Skaggs (1984):
⎧ S (h)
⎫
−γ
⎪ ∫ h(θ ) dθ ⎪
⎪
⎪
K (h) = K sat S (h) l ⎨ 10
⎬
⎪ h(θ ) −γ dθ ⎪
⎪ ∫0
⎪
⎩
⎭
δ
(2-4)
where S is the soil saturation, S = (θ-θr)/(θs-θr), l, γ, and δ- parameters. When water
retention is simulated with the van Genuchten equation (2-2), this equation with γ=1 and
α= 2 can be transformed into the van Genuchten-Mualem equation:
[
K = K sat (S ) 1 − (1 − S 1 / m ) m
l
]
2
(2-5)
Pedotransfer functions need to be used to estimate both Ksat and h(θ) in Eq. (2-5) to
evaluate the integral and calculate the unsaturated hydraulic conductivity.
Artificial neural networks (ANN) recently are useful for PTF development. The feedforward multiplayer ANN, the most popular type for PTF development, has built-in
equations (Pachepsky et al., 1996) and is no different in applications than any other
regression equation. However, coefficients of this equation are never reported, and one
typically has to use software with hardwired ANN equations in it. Rosetta is a Windowsbased program that implements artificial neural network results published by Schaap et
al. (1998), Schaap and Leij (1998), and Schaap and Leij (2000) and is available from
http://www.ussl.ars.usda.gov/models/rosetta/rosetta.HTM. The program implements five
PTFs in so-called hierarchical approach (H1, see also Schaap et al, 2001). This approach
was chosen to maximize the accuracy of the PTF estimates given the particular data
availability. All models in Rosetta estimate saturated hydraulic conductivity and van
Genuchten (1980) retention and unsaturated hydraulic conductivity parameters. Through
use of the Bootstrap Method, Rosetta is also able to estimate the uncertainty of the
estimated hydraulic parameters. Minasny and McBratney (2002) developed the
Neuropack software package (http://www.usyd.edu.au/su/agric/acpa/neuropack
/neuropack.htm). This package differs from the previously discussed software since it is
primarily intended to develop PTFs using neural network-based techniques with the usersupplied. No previously calibrated PTF is shipped with Neuropack except some sample
data.
31
Table 2-6. Parameters of the Ahuja equation (2-3) to estimate saturated soil hydraulic
conductivity
Source
B, 10-3 m s-1
n
Ahuja et al., 1989
2.9
3.21
Franzmeier, 1991
5.0
3.25
Messing, 1989
1.5 – 9.5
1.8 – 3.0
Timlin et al. 1996
2.1
3.29
Rawls et al., 1992
2.8
4.0
32
In general, ANNs provide an excellent flexibility in mapping complex 'input-output'
dependencies (Hecht-Nielsen, 1990). The use of ANNs has, however, some
disadvantages compared with regressions. In particular, the equations built during ANN
training are “opaque,” and ANNs do not distinguish inputs by their significance leaving
the responsibility to select significant inputs to the user.
Estimating solute transport parameters remains a weak point of the PTF development.
Although a progress has been made in estimating solute dispersivity for small soil cores
(Griffioen et al., 1998; Oliver and Smettem, 2003; Minasny and Perfect, 2004), no
methodology exists for such estimates at the pedon scale and at coarser scales.
2.2.7.2. Performance of pedotransfer functions
Performance of pedotransfer functions (PTF) can be evaluated in terms of their accuracy,
reliability, and utility. In broad terms, the accuracy of a PTF can be defined as a
correspondence between measured and estimated data for the data set from which a PTF
has been developed. The reliability of PTF can be viewed as a correspondence between
measured and estimated data for the data set(s) other than the one used to develop a PTF.
Finally, the utility of PTF in modeling can be assessed in terms of the correspondence
between measured and simulated environmental variables.
PTF accuracy has been characterized using various quantitative measures, such as the
mean error, the standard deviation of the mean error, the mean squared error, the
determination coefficient R2, etc. With the same measure, accuracy of existing PTFs
varies appreciably. Pachepsky et al. (1999) randomly sampled literature on PTF and
found that soil basic parameters typically account for 80 - 90 % of the moisture content
variance at capillary pressures of 330 and 15000 cm. The RMSE ranges from 0.02 to 0.07
m3 m-3 at both levels of capillary pressure. The water content at 330 cm ranges from 0.04
to 0.50 m3 m-3 and the water content at 15000 cm ranges from 0.02 to 0.35 m3 m-3 for the
most of soils. With such ranges, the PTF RMSE between 0.02 to 0.07 m3 m-3 is sufficient
in some cases, while in other cases it is unsatisfactory, and improvements in PTF
accuracy are desirable. The hydraulic conductivity PTFs usually have the accuracy about
half order of magnitude.
The reliability of PTFs can be estimated by re-sampling with the data set used in the PTF
development, or by using an independent data set. The re-sampling methods generate
numerous artificial data from the original experimental data and evaluate statistical
properties of interest from its observed variability over all the generated artificial datasets
(Good, 1999). Two methods, the bootstrap (Schaap, 2004) and jackknife (Pachepsky and
Rawls, 1999) have been used so far. The re-sampling methods do not render any
information about the reliability of PTF if it is applied in the region other than that the
development data set has been collected at.
PTFs developed from regional databases produce quite adequate results in the regions
with similar soil and landscape history. Water retention PTFs developed in Belgium
(Vereecken et al., 1989) were the most accurate as compared with 13 others for the data
33
base of the Northern Germany (Tietje and Tapkenhinrichs, 1996). Water retention PTFs
developed for the Hungarian Plain (Pachepsky et al., 1982) were applicable for the
Caucasian Piedmont Plain (Nikolaeva et al., 1986). PTFs developed in Australia were
more accurate for the Mississippi Delta as compared with other regional PTFs (Timlin et
al., 1996a). It remains to be explored whether this observation holds for other cases, and
which soil and landscape features have to be similar in two regions to assure the mutual
reliability of the PTFs.
PTFs developed from the USA database by Rawls et al (1982) appear to be more robust
than PTFs developed from the regional databases. When the reliability of the all-USA
PTFs has been compared with the accuracy of several other regional databases, and the
PTFs were ranked by their reliability, the all-USA PTFs usually had one of the highest
rankings (Tietje and Tapkenhinrich, 1993; Kern, 1995; Timlin et al., 1996b). In general, it
is not possible to predict whether a specific PTF will be applicable at a given site or not.
An example of an interregional comparison of PTFs is given in Fig. 2-11 where 21 PTF is
applied to estimate water contents at 33 kPa and 1500 kPa. (Wösten et al., 2001). This
figure illustrates various situations that are encountered in PTF reliability estimations.
The PTF #1 has the CEC of clay fraction as an essential input. An average value of the
CEC/clay ratio=0.5 was applied to all data to compute the data in Fig. 2-11. PTF # 3 is
inaccurate with Oklahoma data, but it is also not particularly accurate with the data used
for the development, R2 being only 0.58. The PTF # 5 is derived from data on 6 soils, and
it fails when tested with the large database. Similarly The PTF # 7 is derived from data on
only 43 soils. The PTF # 6 for the field capacity was developed with actual field data, and
an attempt to equate the field capacity to the water content at 3 kPa was unsuccessful.
PTFs # 2 and # 8 were derived for soils with the clay minerals that are distinctly different
from those in Oklahoma. They tend to overestimate water retention at 15 bars; for the
same reason PTF # 9, PTF # 13, and PTF # 19 tend to underestimate it. PTFs # 18 and 3
developed for tropical soils do not give satisfactory predictions of field capacity, but
seem to perform well at 1500 kPa. Using the texture only leads to unsatisfactory results at
low water contents (PTF # 17). Overall, PTFs # 11 and # 16 are the best.
The reliability of PTFs may be limited by differences in methods used to measure soil
hydraulic properties that may have profound effect on results (Marion et al., 1994). One
of the method-related factors affecting the PTF development is the support size as shown
in Figures 2-7 and 2-8. Large data bases assembled from different sources to develop PTF
often lose in their quality because of the diversity in measurement methods. There also
may exist a limit in accuracy and reliability of PTFs caused by temporal variation of soil
properties related to the changes in vegetation and soil management. Pachepsky et al.
(1992) have reported 20% changes in water retention at -1 kPa and 5% changes in water
retention at -30 kPa in soils under wheat during growing season in three different climatic
zones. The amount of published data on temporal variations of soils hydraulic properties
remains insufficient. The temporal component assessment may be required in PTFs to
improve their reliability.
34
Estimated volumetric water content (m3m-3)
1
2
3
4
5
6
7
0.5
0.4
0.3
0.2
0.1
0.0
0.0 0.1 0.2 0.3 0.4 0.5 0.0 0.1 0.2 0.3 0.4 0.5 0.0 0.1 0.2 0.3 0.4 0.5 0.0 0.1 0.2 0.3 0.4 0.5 0.0 0.1 0.2 0.3 0.4 0.5 0.0 0.1 0.2 0.3 0.4 0.5 0.0 0.1 0.2 0.3 0.4 0.5
8
9
10
11
12
13
14
0.5
0.4
0.3
0.2
0.1
0.0
0.0 0.1 0.2 0.3 0.4 0.5 0.0 0.1 0.2 0.3 0.4 0.5 0.0 0.1 0.2 0.3 0.4 0.5 0.0 0.1 0.2 0.3 0.4 0.5 0.0 0.1 0.2 0.3 0.4 0.5 0.0 0.1 0.2 0.3 0.4 0.5 0.0 0.1 0.2 0.3 0.4 0.5
15
16
17
18
19
20
21
0.5
0.4
0.3
0.2
0.1
0.0
0.0 0.1 0.2 0.3 0.4 0.5 0.0 0.1 0.2 0.3 0.4 0.5 0.0 0.1 0.2 0.3 0.4 0.5 0.0 0.1 0.2 0.3 0.4 0.5 0.0 0.1 0.2 0.3 0.4 0.5 0.0 0.1 0.2 0.3 0.4 0.5 0.0 0.1 0.2 0.3 0.4 0.5
Measured volumetric water content (m3m-3)
Figure 2-11. Estimating water contents at 33 kPa (○) and at 1500 kPa (●) with 21 PTFs developed in various region of the world
Experimental data on soil water retention and soil basic properties have been extracted from the US Cooperative Soil Survey database
for the state of Oklahoma. From (Wösten et al., 2001). With permission, Elsevier.
35
The reliability of a PTF is not directly related to its utility (Pachepsky et al., 1999). The
latter is affected by sensitivity of the model to PTF predictions, and also by the
uncertainty in other model inputs (Leenhardt et al., 1995). When the PTF uncertainty is
factored in the modeling effort, variation in predictions of different PTFs has to be
considered along with uncertainty of individual PTF predictions. As the experience of
calibrating vadose zone models accumulates, more opportunities appear to compare
calibrated and PTF-estimated hydraulic properties (Wang et al., 2003) as well as to use
PTF predictions as initial estimates for model calibration (Jacques et al., 2002). Using
several different PTFs instead of predictions of a single PTF appears to be a viable option
to explore.
2.2.7.3 Emerging pedotransfer techniques
Developments in pedotransfer technologies are mostly associated with using spatially
dense physical information related to the soil cover. Schematic of this research is shown
in Figure 2-12. The use of topographic information is based on the hypothesis that there
may exist some relationship between soil hydraulic properties and topographic variables
(Pachepsky et al., 2001) because (a) the basic properties of soil are known to be related to
landscape position, and (b) soil hydraulic properties are related to soil basic properties,
Romano and Palladino (2002) used the terrain attributes to recalibrate PTF, and soil water
retention exhibited strong dependence on terrain attributes in the work by Rawls and
Pachepsky (2002) with the US National Cooperative Soil Survey data.
The use of hydrogeophysical, remote sensing, and crop yields data is based on a similar
premise. The basic data on soils affect both sensor readings and soil hydraulic properties.
Therefore, some relationship between these dense data and hydraulic properties can be
expected. Smettem et al. (2004) presented an example of using airborne gamma
radiometric sensing to estimate the clay content of surface soils and using a simple
pedotransfer function (Smettem et al. 1999) to convert this information into a spatial
representation for the slope of the Brooks and Corey water retention function. Timlin et
al. (2003) related the yield maps to the soil field capacity.
Using dense auxiliary data in PTFs is an attempt to trade the data quality for data
quantity. Since a dense coverage can be treated as an image, the image analysis
techniques can be used for segmentation and classification, as well as for evaluation of
spatial structure in soil properties. Also, the data assimilation techniques can be suited to
combine the soil survey and sensor information (McLaughlin, 1995). PTFs developed
with auxiliary data are probably highly site-specific and, therefore, useful for the sites
they have been developed for. Nevertheless, the availability of sensor data can make the
development of such PTFs a viable component of supplying parameters of the models of
water flow in vadose zone.
Since the accuracy of any PTF outside of its development dataset is essentially unknown,
the use of an ensemble of PTFs may be considered. Existence of several models that are
developed and tested in one region but perform relatively poorly in other regions is fairly
common in meteorology. Justifying a selection of a single model has become an
36
Hydrogeophysical methods: GPR, EM, ERT
Remote sensing - passive and active IR, gamma-radiometry,
Yield mapping
Distribution of
basic soil and
sediment properties
(texture, organic
matter,
bulk density, …)
in landscapes
Soil water
retention,
soil hydraulic
conductivity
Topographic variables
Figure 2-12. Using the dense auxiliary data to infer soil hydraulic properties.
37
unsolvable problem. The multimodel ensemble prediction method was developed during
the last decade to address this dilemma (Molteni et al, 1996). The basic idea is to use a
weighted average of prediction obtained from all the models in hand. Since there is no
underlying theoretical formalism from which a probability distribution of model
uncertainty can be estimated, some pragmatic approach must be sought. One such an
approach relies on the fact that climate models have been developed somewhat
independently by different research groups. An ensemble comprising such quasiindependent models is referred to as a multi-model ensemble (Palmer et al., 2004).
Ensemble forecasts offer a way of filtering the predictable from the unpredictable through
averaging – the features that are consistent among ensemble members are preserved,
while those that are inconsistent are reduced in amplitude. What is even more important,
the ensemble itself, as a sample from possible forecast outcomes, can be used to estimate
the forecast uncertainty and the likely structure of forecast errors (Hamill et al. 2004).
Recently, the multimodel ensemble methods have begun to be used in subsurface
hydrology. Ye et al. (2004) suggested averaging of spatial variability models in
unsaturated fractured tuff in a situation when standard information criteria provide an
ambiguous ranking of the models that does not justify selecting one of them and
discarding all others. A limited testing outlined in Chapter 3 of this report suggests that
the multimodel ensemble prediction of soil hydraulic properties is a promising technique.
2.3. Model abstraction implementation
2.3.1
Status of model abstraction in flow and transport modeling
The status of model abstraction in flow and transport modeling has been recently
reviewed by Neuman and Wierenga (2003) They indicate that “…a widely practiced
approach has been to simplify complex hydrogeologic systems for modeling purposes in
an ad hoc and subjective manner. It is generally not clear that this approach captures
adequately all aspects of the complexity that have a significant impact on the problem
under consideration. It is also not clear that such models have the space-time resolution
required to render them comparable and compatible with site data. Without this, the
models remain interesting logical constructs, however, may not be not valid
representations of actual site conditions and may yield unreliable results.”
Currently, solid justification for a model selection is rarely provided. This refers to both
the selection of the conceptual flow and transport models, i.e. process descriptions, flow
and transport domain discretization, forcing and initial conditions, and the selection of the
method to estimate the hydraulic properties. Comparison of performance of two or more
models is rarely made. Most model abstraction methods are not comprehensively tested
for modeling of flow and transport in variably saturated soils and sediments. The
following describes a systematic approach to the model abstraction.
38
2.3.2. Model abstraction process
General requirements to the model abstraction process are objectiveness, systematic
implementation, comprehensiveness, and efficiency (Neuman and Wierenga, 2003).
Model abstraction (MA), as defined in this report, is not about developing models, but
rather about modifying or replacing them. The MA process starts with an existing base
model that can be calibrated and used in simulations. The key output of the model id
defined that provides the necessary and sufficient information to decide on issues of
interest.
The base model may need abstraction for one or more of the following reasons:
• The base model includes a complex description of processes that cannot be
observed well and yet need to be calibrated; the calibrated values of parameters of
those processes are very uncertain.
• The base model propagates uncertainty in the initial distributions, parameters, and
forcing in a manner that creates an unacceptable uncertainty of the key output.
• The base model produces inexplicable results in terms of the key output.
• The base model requires an unacceptable amount of resources for computations,
data preprocessing, or data post-processing, e.g. the base model is not suitable be
used as a part of an operational modeling system that requires real-time data
processing.
• The base model lacks transparency to be explicable and believable to the users of
the key output.
The model abstraction process includes the following steps.
1. Justify the need for the model abstraction
2. Review the context of the modeling problem
3. Select applicable MA techniques.
4. Determine MA directions and simplify model in each direction.
The following provides additional information on each of the steps.
2.3.2.1 Justify the need for model abstraction
Any model simplification requires calibration of the abstracted simpler model and its
confirmation with multiple model runs. Therefore, it is a separate modeling project that
also demands resources, and the need in model abstraction has to be justified.
The lack of transparency of the base model is the reason for abstraction that arises from
the perception of potential users or critics of simulation results. The justification for
abstraction in such case is outside of the modeling project per se. The incomprehensible
results of simulations with the base model or its unacceptable resource demand of base
model are the reasons that essentially preclude the completion of the modeling project.
The justification for model abstraction is simply the need to finish the project.
39
The need in abstraction may stem also from the uncertainty in calibrated parameter values
or in simulation results. The decision to carry out the model abstraction is based on
statistics of parameter estimates and the key output, as discussed below.
2.3.2.1.1 Need in abstraction because of the limited observation ability
A case study of the effect of observability on the model performance has been recently
presented by Kekkonen and Jakeman (2001). They compared two catchment hydrologic
models. Each model had six parameters. Both models were calibrated with streamflow
data. One model explicitly simulated evapotranspiration and was parameterized in this
part, whereas another model did not contain this process as described directly. The model
without the evapotranspiration submodel provided, in general, more accurate
reproductions of streamflow, even on the independent dataset. This difference was more
pronounced at drier catchments. This result was attributed to the calibration with the
streamflow data that contained relatively little information about evapotranspiration.
Authors concluded that the more process complexity one wanted to include in the model
structure, the more types of data and larger information content were required to estimate
the process parameters and to test the model performance. When only rainfall-runoff data
were available, it was difficult to justify a substantial conceptualization of
evapotranspiration.
The model calibration gives the necessary quantitative information to decide on the
reliability of the parameter estimates with respect to their importance for the key output.
A parameter estimate is acceptable if either the uncertainty of this parameter is adequate
as compared with estimates of other parameters, or it is unimportant for predictions of
interest. A diagnostic table (Table 2-7) has been developed based upon the classification
proposed by Hill (1998). In Table 2-7, diagnosis is based on (1) statistical characteristics
of the parameter, and (2) the model prediction uncertainty. In this analysis, the basic
diagnostic measures are: (1) the composite scaled sensitivity of the parameter; (2)
coefficient of variation of the parameter; (3) the scale sensitivity of prediction; (4)
correlation coefficients between estimated parameters, and (5) the updated correlations
coefficients of the parameter. These statistics are directly available or can be computed
according to Hill (1998) from the output of calculations with the available from the most
widely used universal parameter estimation codes, such as UCODE
(www.mines.edu/igwmc/freeware/ucode/) and PEST (www.parameterestimation.com/html/pest_overview.com), as well as from the inverse modeling with
HYDRUS (Simunek et al, 1999). The diagnostic table 2-7 suggests using the model
abstraction where the reliability of calibrated parameters is low. Model structure
abstraction or model parameter abstraction is suggested depending on whether the
parameters are influential or not.
In variably saturated porous media, the complexity of process description related to the
specific parameter or to the group of parameters can preclude a reliable estimation of
these parameters with the additional data from available sources. For example, the
reliable estimation of parameters of the mobile-immobile zone solute transport model has
been found impossible in some field studies (Koch and Flühler, 1993; Field and Pinsky,
40
2000). However, if some additional experiments can potentially provide the information
to improve the parameter estimates, the decision to carry out the model abstraction should
be postponed until the data from such experiments are available. An (exaggerated)
example is the soil water flow model that includes preferential flow but has been
41
Table 2-7. Potential need for model abstraction from the parameter uncertainty standpoint
based on Hill (1998) analysis.
Importance of the parameter Reliability of the parameter estimate
to the key output
Precision of the parameter estimate
Good precision: small CPoor precision:
¶
scaled sensitivity, small
large C-scaled sensitivity ,
large coefficient of
coefficient of variation,
variation, large confidence
small confidence interval
interval
Not important: small PAbstraction of parameter
No indications to use model
scaled sensitivitydetermination may be useful abstraction
Important: large P- scaled
Abstraction of the model
No indications to use model
sensitivity
structure may be useful
abstraction
Uniqueness of the parameter estimate
Poor: the absolute values of Good: all of the Ccorrelation coefficients for
C-correlation coefficients§
of this parameter with some this parameter have
other parameters are close
absolute values less than
to one
0.95
No indications to use model
Not important: the absolute Abstraction of parameter
determination may be useful abstraction
values of P-correlation
coefficients* of this
parameter with the same
parameters as here remain
close to one.
Important: the absolute
Abstraction of the model
No indications to use model
values of P-correlation
structure may be useful
abstraction
coefficients of this
parameter with the same
parameters as here are low
¶
The C-scaled sensitivity (composite scaled sensitivity in Hill, 1998) shows the effect of
the small increment in the parameter value on the simulation of observed values used in
calibration. The large composite scaled sensitivity means that that the simulation of
observed values is very sensitive to the parameter value.
The P-scaled sensitivity (predicted scaled sensitivity) shows the effect of the small
increment in the parameter value on the simulated of key output. The large predicted
scaled sensitivity means that the key output is very sensitive to the parameter value.
§
The C-correlation coefficient of two parameters is computed as the parameter correlation
coefficient (Hill, 1998) when only simulated observations are used. Large C-correlation
coefficient of two parameters means that they cannot be estimated independently from
available observations, i.e. cannot be uniquely defined.
*
The P-correlation coefficient of two parameters is computed as the parameter correlation
coefficient (Hill, 1998) when both simulated observations and the key output are used.
42
Large P-correlation coefficient means that the correlation in estimated values of
parameters, albeit high, does not affect the key output. Small P-correlation coefficient
means non-uniqueness of parameter determination affects the key output.
Equations to compute the C-scaled sensitivity, the P-scaled sensitivity, the C-correlation
coefficients, the P-correlation coefficients, the coefficients of variation of parameters, and
the confidence intervals of parameters are provided by M. Hill (1998).
43
calibrated with the data on soil water regime observed during the period without
substantial rainfall events. No reliable parameters for the preferential flow submodel can
be determined in this case. Similarly, an attempt to find a reliable value of the molecular
diffusion coefficient from a breakthrough experiment fails if the convective transport and
the hydrodynamic dispersion dominate the transport. For the above example with the
preferential flow, soil water regime and concentrations of a conservative tracer can be
monitored during large rainfall events, then the data for parameterization the preferential
flow submodel become available. The decision on abstracting the base model depends on
whether these data will be sufficient or not to determine the reliable parameter values.
A detailed set of guidelines for effective model calibration has been developed for
groundwater modeling (Hill, 2004). But in flow and transport modeling in variably
saturated soils and sediments, the well-tested guidelines for calibration have not been
developed so far. However, the general concepts of the groundwater modeling guidelines
appear to be applicable to the models with nonlinearities encountered in unsaturated zone
modeling. These concepts are:
using the prior information; this means either using the prior probability distributions
of the sought parameter values, or using the measurements of auxiliary values that are not
directly simulated but can be computed from the simulation results. In the former case,
the maximum likelihood method renders for the minimization the composite sum of the
squared model errors and squared differences between prior and posterior parameter
estimates (Neuman and Wierenga, 2003). In the latter case, the composite sum to
minimize includes the squared model errors and squared differences between measured
and computed from simulations auxiliary values (Hill, 1998).
using the measurement errors; this often means assigning weights to the model errors
that are inversely proportional to the error in the data;
Since the calibration problem is mathematically ill-posed, there exists a probability of
finding a set of parameters that do not provide the global minimization of the minimized
sum of squared differences. This probability generally increases when the data accuracy
is lower. To mitigate this source of error, it is advised to use a preliminary scanning of the
plausible parameter value domain as it is done in the SUFI algorithm (Abbaspour et al.,
1997; Schmied et al., 2000). Such scanning allows one to define sub-domains in the
parameter space where the correct set of parameters is located with the highest
probability.
The prior information about flow and transport parameters has to be obtained either from
the measurements at similar sites or from the pedotransfer functions. It is strongly
advised to use an ensemble of pedotransfer functions developed with different,
sufficiently large databases (Appendix A).
2.3.2.1.2 Need in abstraction because of uncertainty in key model outputs
The system behavior has to be predicted under realistic forcing different from the forcing
used in the model calibration. Monte Carlo simulations have to be set with correlations
between parameters utilized in scenario generation. If the Gaussian distribution can be
44
used for original or transformed parameters, the method suggested by Clifton and
Neuman (1982) should be used to produce random realizations. If the correlations
between inputs are known, the uncertainty is substantially reduced (Meyer et al., 1997).
Pohlman et al. (2002) modeled density-driven flow and radionuclide transport at the
underground nuclear test site and showed that the incorporation of correlation between
hydraulic conductivity and recharge significantly reduced the uncertainty in both travel
time and transverse plume location. The climate data generator CLIMGEN
(http://www.bsyse.wsu.edu/climgen) can be used to generate the weather scenarios.
The uncertainty of the key output has to be quantified using probability distribution
function to see the expected range of predictions given the uncertainty of inputs.
If this range is unacceptable for the purposes of performance assessment, i.e. the median
predicted value cannot be statistically significantly discerned from the critical value of
the key output, then the model abstraction can be considered. Parameters and inputs of
the abstracted model may be estimated with lower uncertainty, and the output may
become more reliable.
2.3.2.2 Review the context of the modeling problem
The context of the modeling problem has to be reviewed to assure the objectiveness and
the comprehensiveness of the model abstraction. It needs to be realized what details and
features of the problem are omitted or de-emphasized when the abstraction is performed.
The context of the modeling problem has been defined by Neuman and Wierenga (2003).
They list the following issues that constitute the context:
1. What is (are) the question(s) that the base and abstracted models are to address?
The answer should consider (a) the potential or existing problem in which
modeling is one of the solution instruments, (b) the potential or existing causes of
the problem, (c) the issues needing resolution, and (d) the criteria to decide on
efficiency of the resolution. The key output has to be provided with the spatial and
temporal scale at which it is evaluated. Acceptable accuracy and uncertainty of
the model output have to be established from the end users. In some cases there
are mandatory regulations on performance measures that articulate the statistics to
use in the particular case. If there are no such regulations, then the statistics have
to be selected and defined; it should describe simple and clear ideas about the
correspondence between the data and simulations, e.g., how the variability of
model errors compares with the variability in the data; do the model residuals
have trends; is there a systematic relative or absolute error in predictions, etc.
2. What kind of data is available to calibrate the base and the abstracted models and
to test them with respect to the key output? The essential condition is to have the
database as broad as possible. It has to include the data from public and private
sources, cover both quantitative and qualitative (expert) information, and
encompass both site-specific and generic information. The list of the base model
inputs and outputs provides a convenient template for the necessary part of the
45
database. It is imperative to have statistics of all available model inputs and
measurable model outputs, including (a) the initial distributions of water contents
and/or matric potentials and concentrations of solutes of interest, (b) the pore
space and surface properties including horizon or layer thickness, porosity, bulk
density, adsorption parameters, and some cases redox conditions (c) the forcing
that provides the boundary conditions and the source/sink terms, and (d) the
model parameters typical for the site. The latter can be inferred from the ensemble
of pedotransfer functions provided the basic soil survey and/or borehole log data
are available to serve as the pedotransfer inputs. Statistics includes the type of
statistical distribution, the median values and variability measures, the
information about observed outliers, and the correlations between parameter
values.
3. To make sure that the abstracted models are sound, the additional information has
to be collected insuring that the abstracted models include descriptions of all
essential processes of flow and transport for given site. This information may be
of lower quality compared with the necessary part of the database. However, one
has to be sure that some small-scale internal heterogeneities will not have a
dominant effect on flow and/or transport at the scale of interest. It is easy to
become right for a wrong reason without the information on possible effects of
processes that are not included in the model. To become wrong for a right reason
is much more responsible way in this case.
2.3.2.3 Define applicable MA techniques
MA can lead to simplifications via
• the number of processes being considered explicitly,
• process descriptions,
• coarsening spatial and temporal support
• the number of measurements for the reliable parameter estimation,
• reduced computational burden,
• data pre-processing and post-processing.
The number of processes can be decreased using abstractions with hierarchies of models,
and delimiting input space. Using the hierarchy of models entails a smaller number of
processes accounted for in the model. Delimiting input space can be done by changing
from two-dimensional to one-dimensional representation of flow and transport in variably
saturated soils and sediments. Not only this procedure decreases a pre-processing time,
demand on the initial distribution data, and a runtime, but it also excludes the need in
defining components of transport coefficient tensors that require elaborate measurements
to be found by calibrations. If the flow is predominantly vertical, eliminating the second
dimension in many cases causes a relatively small change in flow and transport
description in variably saturated soils
Process descriptions can be simplified using abstractions with a hierarchy of models,
limiting input domain, and scale change. As the complexity of the model porous media in
46
Fig. 2-2 decreases, both flow and transport can be described by simpler models.
Upscaling can generate models for coarser scales. These models can still be applicable
for computing the measure of the performance. At the same time, they have a parameter
set that is much smaller and simpler. The same is true in case of applying aggregation.
However, the coarser scale parameters are ‘lumped,’ i.e. they represent a joint effect of
several processes and of their variation at the subgrid scale. It is usually impossible to
relate the values of these parameters to the basic soil and landscape properties; therefore,
the prior estimates of their values are not possible.
If the scale is coarsened but the process description is not changed, one should expect the
scale-dependence of parameters. There typically exists a mismatch between the
observation scale and computational grid scale; therefore an extreme caution has to be
exercised when observations are used to render parameter values for simulations. In
unsaturated zone modeling, the reliance on laboratory measured hydraulic properties is
much higher than in saturated zone projects. Coarsening of the scale typically creates
computational grid units that are larger than the laboratory samples. In such case, scaling
abstraction techniques have to be applied and the empirical scaling laws have to be
established to convert the laboratory measurements to the grid-scale parameters.
Model abstraction does not decrease the number of measurements needed to calibrate the
base model. Ultimately, the performance assessment can be conducted with the aid of
highly simplified, abstracted model, and yet the base model has to be calibrated and
tested to assure that it is able to represent the natural complexity of the hydrological
content, and that it can work as a starting point for abstraction. The number and type of
measurements affect both parameter estimation and model performance evaluation. The
measurements have to provide (a) the reliable values of parameters to which the key
model output is sensitive, and (b) the reliable key model output.
The performance of models can be evaluated with a variety of statistics, most of which
normalize the variability of errors by the variability in data (Neuman and Wierenga
2003). High variability of the data may cause the statistic value to be relatively low not
because the model is good but because the data are poor.
It has been often assumed that spatially and temporally averaged values are less prone to
the predictive uncertainty. This is generally true for statistically homogeneous and
stationary systems. Spatial-temporal fields of water contents, concentrations, and soil
water potentials usually do not conform these statistical requirements. Nevertheless, the
change of scale may actually reduce the variability of the data because of the temporal
persistence in differences between measurements in different locations (Grayson and
Western, 1998). If the persistence is documented, the coarse-scale average water
contents, for example, can be estimated with measurements of the lower density.
Reducing computational burden can be achieved using any of the abstractions. Large
runtime can be caused by either the numerical stability requirements or by the temporal
detail in the data used to provide input or to evaluate the performance measures. Strong
nonlinearity of the models for variably saturated water flow imposes substantial
47
limitations on the integration time step, especially for the coarse-textured soils and
sediments. It is always worthwhile to see whether reformulation of boundary conditions
on the soil surface (similar to the one done by van Dam and Feddes (2000) or
reformulations of water transport equations (similar to the one done by Pan and Wierenga
(1995)) may improve the computation speed. Similarly, it is advisable to use look-up
tables to avoid multiple computations of nonlinear constitutive relationships. Coding the
independent submodels that can be executed in parallel, decreases the number of grid
variables that participate in each iteration process present in the model. Wu et al. (2002)
used parallel computing in simulations of unsaturated flow and transport for the Yukka
mountain project; Englert et al. (2004) demonstrated using parallel computing to solve
the saturated-unsaturated flow and transport problem for a large area of pesticide
application. However, the limitations of numerical methods and the need to run multiple
simulations may still call for model abstraction. Scale change via upscaling or
aggregation typically results in much smaller number of grid units and usually in much
faster simulations. The metamodeling requires fairly large number of preliminary model
runs to generate the development and testing datasets, and may be efficient if the number
for future runs of a neural network metamodel is substantially larger than the number of
runs needed to develop this metamodel.
Data pre- and post-processing may consume a substantial part of the resources available
for the modeling project. Most of model abstraction techniques require changes in the
input data used for modeling. The input data are expected to be less voluminous for
simpler models. However, the data may also be less certain, and a larger number of
simulation runs may be needed to account for this uncertainty. In case of abstractions
done by scale change or changing dimensionality, a computational unit, i.e. a grid cell,
needs to be characterized by the porous media properties and by forcing at the
boundaries. Because of the non-linear effects of subgrid heterogeneities on flow and
transport, such averaging is never certain, and such uncertainty has to be simulated on top
levels of the uncertainty of initial water contents and solute concentrations, pore media
properties, and forcing variables at the boundaries at the fine scale.
In case of extremely large numbers of computational units, even the storage of simulated
water content, matric potentials, and concentration fields may become an issue. Using
wavelets for data compression may be needed to overcome this difficulty (Mallat, 1998,
Neupauer and Powell, 2005).
2.3.2.4 Define and execute model abstractions
Model abstraction, in general, results in smaller number of independent parameters to
measure/estimate and/or the lesser amount of computations. The particular procedure of
MA depends on the purpose of abstraction and on the available resources. Each
abstraction is a separate sub-project that requires justification, planning, and milestonemetered execution. The feasibility of a particular model abstraction can be judged on the
basis of resources spent cumulatively on five stages of the abstracted model application,
namely (1) pre-processing, (2) calibration, (3) simulation, (4) post-processing, and (5)
reporting. For the specific model abstraction procedure, it is necessary to make sure that
48
the software for using the abstracted model is available, that the output of the abstracted
model will be compatible with the available data for calibration and testing purposes, that
the pre- and post processing tools are in place. The necessary steps are also to estimate
the time and computational resources gain from the abstraction, as well as the personnel
available to perform and explain the abstraction. The abstracted models need to be
calibrated as described in 2.3.2.1.1. Plausible scenarios described in 2.3.2.1.2 allows one
to confirm that the abstraction has alleviated or eliminated the problem encountered with
the base model, that it did not create any similar or new problems.
A model that is the result of abstraction can be abstracted further if necessary. Such
sequential abstraction has to be done in a systematic manner to assure the understanding
of the role of each abstraction step in the performance of the final model.
2.4 Concluding remarks
Very complex models were developed as the results of advancements in understanding
and quantifying environmental and social phenomena, as well as industrial processes.
The complexity of models is caused, in particular, by the large number of simulated
processes, large number of interacting entities, differences in scales at which the
interactions occur, large number of feedbacks to simulate. Model abstraction, understood
as model simplification, is the field of active research in a multitude of scientific and
engineering fields. The more actively models are used in a field the more research on
model abstraction can be found. Each research or engineering field has specific types of
models in use, and the model simplification techniques are field-dependent. The intensive
use of models in subsurface contaminant hydrology has resulted in development of many
model abstraction techniques. They have been briefly summarized in previous sections.
Model abstraction in contaminant hydrology has been used mostly in research. Potential
benefits of model abstraction include improvement of understanding and communication
of modeling results, more robust predictions, and better understanding of essential factors
and their representation in models. This makes model abstraction an attractive
methodology for engineering modeling applications. The model abstraction process, as
described in previous section, is a transparent step-by-step formalized procedure of
justification of the use of a simplified model. It is demonstrated in the Chapter 3 of this
report.
Engineering applications of models abstraction require model calibration and uncertainty
analysis toolboxes and appropriate user interfaces. Such toolboxes and interfaces are
currently being developed by several agencies (Castleton et al, 2002; Fine et al., 2002).
Availability of the software tools will greatly simplify the model abstraction use.
An important feature of models abstraction is the explicit treatment of model structure
uncertainty. The model structure, along with the data uncertainty, and scenario
uncertainty, is known to introduce the uncertainty in modeling results. Unlike the
uncertainty in input data, in model parameters, and in scenarios, the effect of the model
structure uncertainty on the uncertainty in simulation results is usually impossible to
49
quantify in statistical terms without making very strong assumptions. With model
abstraction, a series of models with feasible structures can be built and evaluated in a
systematic manner. Each of the models is evaluated from results of an ensemble of
simulations by its accuracy to measurement data and by its predictions with respect to
scenarios that have not been observed. Model abstraction is always performed in the
uncertainty context.
It is observed in diverse modeling projects that several models may have different
structures but the same accuracy with respect to data of measurements. As it is not
possible to decide which model is “better” or “more correct, ” the idea of using an
ensemble of models becomes more and more accepted. Model abstraction can generate
models of different structures in a systematic way, thus creating a model ensemble.
It is expected that as the model abstraction methodology will develop, knowing about
model abstraction and applying it to the regulated sites will be helpful for the NRC
licensing staff. Reviewers will have a means to determine whether the models submitted
to support licensing actions adequately represent the site, and whether the investigations
adequately represent important features, events, and processes for the site. Managers will
have an advisory tool to decide whether the requests for additional information are
targeted tat sensitive site parameters and processes. On the other hand, the NRC licensees
will have a device to determine whether a simple model that is easy to understand and
communicate to regulators, stakeholders, and the general public can adequately represent
their site.
50
3 MODEL ABSTRACTION CASE STUDY
3.1 Introduction
Selection of abstraction techniques and their combinations must take into account the
uniqueness of the environmental system being modeled. This relates to the concept of
“uniqueness of place in environmental modeling” (Beven, 2002; Neuman and Wierenga,
2003). After reviewing numerous available experimental data sets to develop a model
abstraction test case study, we selected a humid site that is relevant to NRC
decommissioning site reviews.
The first and main objective of the test case was to understand how model abstraction can
affect performance assessment of contaminant migration at a relatively humid site where
transport may be affected by the presence of soil macropores and related preferential flow
phenomena. Specifically, the test case is designed to assess the capabilities, limitations, and
usefulness of selected model abstraction methods.
3.2 Description of experiments
3.2.1 Field setup
The experimental field is located at Bekkevoort, Belgium. It is situated at the bottom of a
gentle slope and is covered with a meadow. The soil is classified as Eutric Regosol (FAO,
1975). Typically, the top 1 meter shows three soil horizons: an Ap horizon between 0 and
25 cm, a Cl-horizon between 25 and 55 cm and a C2 horizon between 55 and 100 cm (Fig.
3-1).
A trench, 1.2 m deep and 8 m long, was dug at the field site. The site-specific soil
description was performed along one side of the trench. The Ap horizon thickness varied
along the trench between 8 and 15 cm depth. From a horizontal distance of 1 m from the
side in the trench, a 30 cm thick, sandier layer was observed beneath the Ap horizon. The
underlying Cl and C2 horizons reached depths between 75 -90 cm. Beneath the C2, a Bt
horizon was found. Between 1.5 and 2.2 m from the side of the trench, a 10 cm thick sandy
inclusion at a depth of 75 cm was observed. Texture analyses were done at several locations
in the trench (Table 3-1). A study of water retention along a 30-m trench in the same soil at
the adjacent site (Mallants et al., 1996) was performed. Values of van Genuchten
parameters were found from water retention measurements done at 5-cm long and 5.1-cm
diameter cores with sand-box apparatus for capillary pressures of 1, 5, 10, 50, and 100 cm,
and with pressure cell for capillary pressures of 200, 630, 2500, and 15,000 cm.
The grass cover was removed from the experimental area. A plastic sheet covered one side
of the trench to isolate the disturbed trench zone. The trench has been instrumented as
shown in Fig. 3-2. Volumetric water content was measured with TDR. Sixty two-rod TDR
probes (25 cm long, 0.5 cm rod diameter, 2.5 cm rod spacing) were installed along the
trench of 5.5 meter at 12 locations with the 50-cm horizontal spacing at five depths of 15,
51
Figure 3-1. Soil profile at the Bekkevoort experimental site (after Jacques, 2000). With
permission.
52
Table 3-1. Average soil texture (%) along the trench at the Bekkevoort site.
Depth, cm No. samples >50 µm 50-20 µm 20-10 µm 10-2 µm
15
7
58.6
19.3
6.4
4.5
35
8
56.7
18.9
7.8
3.2
55
5
57.3
17.6
6.6
3.7
75
3
49.6
21.2
7.9
4.4
95
4
43.8
30.03
7.4
4.5
53
<2 µm
11.1
13.3
14.8
17.4
14
Figure 3-2. View of the instrumented trench (after Jacques, 2000). With permission.
54
Figure 3-3. Locations and numbering of TDR-probes (z), tensiometers ({) and
temperature probes (•). Patterned rectangles show Ap, C1 and C2 horizons. Dashed lines
show the average position of the horizon boundaries. Average values of clay, silt, and sand
content are given for the probe installation depths. The numbers in pentagons show the
numbering of locations monitored by the five-probe profiling. From (Pachepsky et al.,
2005). With permission, SSSA.
55
35, 55, 75, and 95 cm (Fig. 3-3). Pressure head was measured with tensiometers.
Tensiometric porous cups (6 mm diameter, 25 mm long) were installed at a horizontal
distance of 10 cm from each of the 60 TDR-probes.
Temperature probes (6 cm long) were installed at 5 depths on distances of 1.3, 2.7 and 4.2
m from the trench.
The TDR-measurements were done with a Tectronix 1502B cable tester. The automated
system of Heimovaara and Bouten (1990) was used to control, retrieve, store and analyze
the measurements of the travel time of an electromagnetic wave along the TDR - and the
soil impedance. Apparent dielectric constants were related to the water content via the sitespecific calibration curve (Jacques et al., 1999). One measurement cycle for all TDRprobes took approximately 35 min. The porous cups were connected with water-filled
capillary tubes to pressure transducers. These transducers were incorporated in an
electronic circuit (SDEC France, Reignac/Indre). Automated measurements of the pressure
transducers were controlled and stored by a Campbell CR10X datalogger and AM416
multiplexers (Campbell Scientific Ltd, Leicestershire, UK). The datalogger and
multiplexing system also controlled the soil temperature measurements.
Passive capillary samplers (PCAPS) were installed at two 15 and 55 cm depths at the
distance of about 5 m from the trench. As discussed by Holder et al. (1991), PCAPS are
able to measure water and solute fluxes under both saturated and unsaturated flow
conditions by imposing a hanging water column to the soil. After combustion at 400 °C
during 8 h (Knutson et al., 1993), the unraveled wicks (Amatex 3/8-inch high density, no.1
0-864KR-02) were mounted in a symmetric way on a 10 mm thick PVC plate. Each wick
covered an area of 17 x 17 cm2. Three wick samplers were put together on one PVC plate.
In the middle of each unit (i.e. the 17 x 17 cm2-area), a hole was drilled from where the
wick was connected to a collection chamber through a flexible PVC tube. A nylon mesh
was put on the top of the wick filaments to prevent the penetration of soil particles into the
PCAPS (Fig. 3-4). For installation in the soil, an excavation of 60 cm width, 10 cm height
and 60 cm depth was dug into the trench wall (Fig. 3-5). The top surface of the excavation
was carefully levelled. The base plate containing three PCAPS-units was put into the
excavation and pushed against the top by hammering wooden blocks under the PCAPSunit. The flexible tube and collection chambers were then placed at the bottom of the trench
in such a way that the distance between the base plate and the bottom of the wick was 90
cm A sampling tube connected the collection chamber with the soil surface. To collect
water from the collection chamber, a capillary tube was put into the sampling tube, and
suction was applied at the soil surface.
Rainfall was measured and recorded near the trench at a plot with a catch area of 200 cm2.
Cumulative rainfall was written continuously with a floated pen system on a paper (0.1 mm
precipitation intervals) fixed on a drum, rotating with a speed of 1 cm per hour.
After all devices were installed, the trench was filled. A thin layer of gravel (1 to 2 cm) was
evenly distributed on the study area: (i) to decrease the erosive effect of the rain impact on
the bare soil surface, (ii) to minimize the evaporation from the soil surface, and (iii) to
decrease the growth of weed on the experimental plot.
56
Figure 3-4. Three passive capillary samplers on the PVC plate (after Jacques, 2000). With
permission.
57
Figure 3-5. Installation of passive capillary samplers (after Jacques, 2000). With
permission.
.
58
Field measurements started on March 11, 1998 (day 0) and finished on March 31,1999 (day
384). Weeds were regularly removed from the site during the summer. The tracer
(CaCl2⋅H2O)
was manually applied on August 28, 1998 (day 168). On an area of 8 x 2 m2, a total of 6000
g was applied using a water depth of 0.5 cm resulting in a concentration of 75 g⋅l-1. The
experimental area was subdivided in 16 squares of 1 m2 and solute was applied in three
steps (2 times 1.5 liter on each square and then 1 liter on 0.5 m2). The resident solute
concentration was calculated from TDR measurements using calibration curves for each
soil horizon.
Frequencies and types of the monitoring measurements are summarized in the Table 3-2.
Total of 400,000 measurements were accumulated during the monitoring period.
3.2.2 Dataset overview
3.2.2.1 Precipitation
Fig. 3-6 shows the hourly rainfall rates and the cumulative precipitation during the
experiment. A total rainfall amount of 109 cm was recorded during the period of the
experiment. The average rainfall in this region is around 80 cm. The amount of rainfall was
rather exceptional since it was the largest amount of rainfall during a 65-year period
recorded at an official nearby weather station (Ukkel, Brusssels). In the time period of the
experiment, the pluviograph in Ukkel had recorded 106 cm.
On days 185 and 186 (13 and 14 September 1998), an extreme rainfall event occurred
which resulted in large scale flooding in the region of the experiment. At the experimental
site itself, no ponding was observed during the storm. However, flooding occurred at the
foot of the hill.
After the storm, there was one more dry period, and an almost linear increase in cumulative
precipitation was observed beginning from the day 237.
3.2.2.2 Soil water contents, capillary pressures and soil temperatures
Time series of average at each depth water contents and capillary pressures are shown in
Fig. 3-7 and Fig. 3-8. Measured average water content varied in time in a limited range
between 0.3 and 0.4 due to the absence of water uptake by plants, limited evaporation and
the relatively high continual rainfall during the experiment. As could be expected, the
temporal water content variations were more pronounced in the topsoil than at larder
depths.
Averaged along the trench capillary pressures varied in time in a relatively small range
(Fig. 3-8). The largest temporal variation was observed during the first 200 days of the
experiment where periods of heavy rainfall alternated with long dry periods. During the last
180 days of the experiment, the pressure heads showed a fast oscillatory behavior between
e.g. 0 and -50 cm at the 15 cm depth and 0 and -20 cm at the 55 cm depth due to the typical
rain pattern during the last 100 days.
59
Table 3-2. Types and frequencies of the monitoring measurements.
Quantity
Water content
Resident solute concentration
Pressure head
Temperature
Water fluxes
Solute fluxes
Rainfall
Device
TDR-probes
TDR-probes
Tensiometers
Temperature probes
Passive Cap. Lys
Passive Cap. Lys
Pluviograph
60
Frequency
Every 2 hr
Every 2 hr
Every hr
Every hr
Every 2-3 days
Every 2-3 days
Continuously
Positions
5 d x 12p
5 d x 12 p
5 d x 12 p
5dx3p
2dx3p
2dx3p
1
120
Cumulative rainfall (cm)
100
80
60
40
Daily
precipitation (cm)
20
0
8
6
4
2
0
0
50
100
150
200
250
300
350
Time (day)
Figure 3-6. Cumulative rainfall and daily precipitation during the experiment.
61
400
0.42
Depth 15 cm
0.38
0.34
0.30
0.26
Depth 35 cm
0.38
0.36
0.34
0.32
Water content(cm3/cm3)
0.30
0.28
Depth 55 cm
0.38
0.36
0.34
0.32
0.30
Depth 75 cm
0.40
0.38
0.36
0.34
0.32
Depth 95 cm
0.40
0.38
0.36
0.34
0.32
0.30
0
50
100
150
200
250
300
350
Time (day)
Figure 3-7. Time series of average water contents at each of the measurement depths. Solid
lines correspond to the measured values at 50%, dotted lines at 25% and 75% probability
levels.
62
400
0
-50
-100
-150
Depth 15 cm
0
-20
-40
-60
-80
Pressure head (cm)
Depth 35 cm
0
-10
-20
-30
-40
Depth 55 cm
0
-10
-20
-30
Depth 75 cm
20
0
-20
Depth 95 cm
-40
0
50
100
150
200
250
300
350
Time (day)
Figure 3-8. Time series of average capillary pressures at each of the measurement depths.
63
400
The lateral and temporal variability of soil water contents are characterized in Tables 3-3
and 3-4. Coefficient of variation (CV) of water content in time was in the range from 0.6%
to 8.1% (Table 3-4). Mean value of CV decreased with depth from 4.7% to 2.0%. Spatial
variations in water content along the trench were more pronounced as compared with
temporal variations. The CV values were in the range from 0.4% to 15.5%, which was
wider than the range of temporal variability. Mean value of CV for space water content
variability did not have depth-related trend. Rainfall was distributed evenly at all locations,
and variability in soil water content has been likely a result of soil heterogeneity and
hydraulic properties variability.
Two-dimensional spatial distributions of water content for several observation days are
shown in Fig. 3-9. There is a definite structure in the data, so that some parts of the profile
remain consistently wet, and other parts stay relatively dry.
Time series of averaged along the trench soil temperature are shown in Fig. 3-10. Values of
temperature had a typical annual sine wave trend and were in the range between 0.4 and
25.6C° on 25cm depth and in the range between 4.0 and 18.3C° on 95cm depth. The
average annual temperature decreased slightly from 11.0 to 10.8C° with depth. The spatial
variability of temperature was very small at all depths. The coefficient of variations of
temperature had no depth-related trend and was in the range from 1% and 0 to 6%.
3.2.2.3 Soil water fluxes
Cumulative soil water fluxes measured with passive capillary lysimeters are shown in Fig.
3-11. The lysimeters responded to the large precipitation events. The variability in values of
the fluxes was substantial and grew as the experiment progressed.
3.2.2.4 Laboratory data on soil water retention
Parameters of the van Genuchten equation (2-2) for the samples taken along 30-m trench
are shown in Fig. 3-12. Statistics of van Genuchten parameters are in the Table 3-5. The
variability of the van Genuchten parameters α and n was relatively high in all three
horizons. There was a weak correlation between saturated water contents θs and lg α. In all
three horizons, and a moderate correlations between lg α and lg n. The same van Genuchten
parameters determined in different horizons did not correlate.
3.3 The base model and the need in model abstraction
The base model was the single continuum model of variably saturated porous media
simulated with the Richards equation. The software to solve the Richards equation and to
estimate its parameters was the HYDRUS-1D (Simunek et al., 1998). HYDRUS-1D is one
64
Table 3-3. Coefficients of variation (%) of soil water contents in time.
Depth
Coefficients of variation
15 cm 35 cm 55 cm 75 cm 95 cm
Min over all probes at a given depth
2.7
1.4
0.7
0.6
1.2
Max over all probes at a given depth
8.1
5.5
3.1
4.6
2.8
Mean over all probes at a given depth
4.7
3
2
2.2
2
Table 3-4. Coefficients of variation (%) of soil water content along the trench
Depth
Coefficients of variation
15 cm 35 cm 55 cm 75 cm 95 cm
Min over all observation times
3.6
3.6
0.4
1.1
2.2
Max over all observation times 11.6
11.2
7.4
9.5
15.5
Mean over all observation times 7.2
6.2
4.3
5.4
7.5
65
230
235
240
245
250
256
257
260
265
Figure 3-9. Consecutive images of two-dimensional distributions of soil moisture contents; days from the beginning of the experiment
are shown.
66
30
Depth 15 cm
25
20
15
10
5
0
Depth 35 cm
20
15
10
5
Temperature Co
0
20
Depth 55 cm
15
10
5
0
Depth 75 cm
15
10
5
0
Depth 95 cm
15
10
5
0
0
50
100
150
200
250
300
350
400
Time (day)
Figure 3-10. Time series of soil temperature. Solid lines correspond to the average
measured values, dotted lines correspond to minimum and maximum values at each depth.
67
200
55 cm
Cumulative flux (cm)
15 cm
150
100
50
PCAPS-1
PCAPS-2
PCAPS-3
Rainfall
0
100
200
300
400
100
200
300
400
Julian day
Figure 3-11. Cumulative rainfall and soil water fluxes measured with passive capillary
lysimeters (PCAPS) at two depths.
68
Ap
C1
C2
0.10
0.08
0.06
0.04
0.02
0.00
0.55
θs
0.50
0.45
0.40
α
0.35
0.30
0.014
0.012
0.010
0.008
0.006
0.004
0.002
0.000
3.5
n
3.0
2.5
2.0
1.5
1.0
0
10
20
30 0
10
20
30 0 5 10 15 20 25 30 35
Distance along the trench (m)
Figure 3-12. Van Genuchten parameters derived from water retention of samples taken
from three soil horizons along the 30-m trench.
69
Table 3-5. Statistics of van Genuchten parameters of water retention in samples along the
30-m trench
θr,Ap θr,C1 θr,C2 θs,Ap θs,C1 θs,C2 lg αAp lg αC1 lg αC2 lg nAp lg nC1 lg nC2
Min
0.000 0.000 0.000 0.348 0.330 0.379 -2.966 -2.947 -2.865 0.141 0.103 0.148
Average 0.040 0.015 0.044 0.411 0.363 0.431 -2.201 -1.951 -2.466 0.236 0.143 0.249
Max
0.083 0.075 0.101 0.484 0.424 0.502 -1.845 -1.536 -2.113 0.552 0.374 0.424
Coefficient
of variation
57.7 135.1 53.8 7.2 5.2 7.7 11.0 13.2 9.7 41.0 34.8 29.9
(%)
Correlation coefficients significant at 0.05 significance level
θr,Ap
1.000 - 0.268 - -0.332 0.349
0.678
1.000 - -0.383 - -0.338 0.851 0.254
θr,C1
θr,C2
1.000 - -0.381 - -0.684 0.834
θs,Ap
1.000 0.430
θs,C1
1.000 0.558
θs,C2
1.000
0.444 0.376 -0.285 -0.329
lg αAp
1.000
- -0.453 - -0.313
lg αC1
1.000
0.282 -0.630 -0.333
lg αC2
1.000
- -0.899
lg nAp
1.000
lg nC1
1.000 0.340
lg nC2
1.000
70
of the most widely used codes for unsaturated flow and solute transport modeling that uses
the finite element method to solve the Richards equation. It has options for nonisothermal
liquid and vapor flow and heat transport. Constitutive relationships include van Genuchten
and Brooks and Corey water retention functions. Information on soil texture can be used
along with pedotransfer functions to determine water retention and hydraulic conductivity
parameters. The code includes inverse modeling capabilities for estimating hydraulic and
transport parameters. The upper boundary condition includes standard constant pressure
and constant flux conditions in addition to meteorological forcing. Options for the lower
boundary condition include unit gradient and seepage face (Scanlon, 2004).
The HYDRUS-1D code was applied to simulate the dataset presented in the sections 3.2.2.1
and 3.2.2.2 (Jacques et al., 2002). These authors calibrated the model using three different
measurement sets in the inverse optimization: the water content data averaged by depth for
each depth θ(t,z); the pressure head data averaged by depth for each depth ψ(t,z); the hourly
rainfall rates that are smaller than 0.5 cm h-1 q(t,0). The upper boundary condition was
defined by hourly rainfall rates. The evaporation was assumed to be negligible during the
experiment due to the presence of a thin layer of gravel of soil surface. The lower boundary
condition was defined as a zero pressure head gradient, i.e. free drainage. The initial
condition was determined by linear interpolation of the measured pressure heads between
the five depths at day 0.
Two approaches were used for the estimation of the soil hydraulic parameters. First,
parameters were estimated using time series of observations for a specific depth assuming a
homogeneous soil profile. Parameters were interpreted as effective parameters
representative for an equivalent homogeneous porous medium between the soil surface and
a particular depth. The objective function included three data sets: q(t,0), θ(t,z) and ψ(t,z)
for the particular depth. Three parameters (θs,α and n) of the van Genuchten equation (2-2)
and two parameters (Ksat and l) of Genuchten-Mualem equation (2-5) were estimated for
each of five observation depths. Second, a layered soil profile with different parameters for
each layer was assumed. The objective function included eleven data sets: q(t,0) and data
sets θ(t,z) and ψ(t,z) for five depths. Twenty soil hydraulic parameters were estimated
within single optimization. Data from passive capillary lysimeters were used neither in
calibration nor in the model performance assessment.
The calibrated model reproduced the time series of soil water contents very well. Simulated
pressure heads decreased less compared to observations at all depths during the dry period
in the summer (day 140-160) and were significantly overpredicted at depth of 35 cm
between days 150 and 200 during the wet period (see Jacques et al., 2002 for details). The
total simulated cumulative infiltration was only from 50.2 to 83.8% of the total amount of
rain (Fig. 3-13). The largest discrepancy was obtained during the heavy rainfall on day 185,
and smaller differences during moderate rainfall conditions (e.g. days 100 and 150, and 250
and 350). The model predicted substantial runoff. However, no runoff or ponding were
observed in the field.
This inexplicable result indicated the need for the model abstraction.
71
Figure 3-13. Measured cumulative rainfall and simulated cumulative actual infiltration for
the different objective functions (Jacques et al., 2002). Φ1, Φ2, Φ3, Φ4 and Φ5 denote to the
parameter optimization at depths of 15, 35, 55, 75 and 95 of homogeneous soil profile; Φall
denotes to the multi-layered soil profile optimization.
72
3.4 Review of the model abstraction context
Model abstraction starts with defining purpose and resources. As discussed in the Chapter
2.3.2.2, the main directions of the context review are:
(1) What is (are) the question(s) that the base and abstracted models are to address? What is the
key output of the model that quantifies those questions? How the results of abstractions will
be evaluated?
(2) What kind of data is available to calibrate the base and the abstracted models and to test
them with respect to the key output?
(3) What is the additional information that has to be collected to make sure that the
abstracted models include descriptions of essential processes of flow and transport for
given site.
3.4.1 The key output
The NRC staff review of performance assessments of nuclear facilities, e.g.,
decommissioning of the facilities, management of low-level and high-level radioactive
waste disposal sites, frequently involves the review of models of water flow and solute
transport in soils. Because water flow is the essential driver of the contaminant transport in
subsurface, an accurate prediction of water fluxes is the precondition of correct estimating
transport of contaminants. Therefore, the soil water fluxes were selected as the key output
of the base and abstracted model.
Inspection of the soil water fluxes measured at depths of 15 and 55 cm with PCAPS
revealed gaps in measurements and equipment malfunctioning. Soil water fluxes
measurements started 40 days after the beginning of water content and pressure head
measurements. Time interval between measurements varied from 1 to 8 days. The
measured cumulative fluxes after rainfall event on the day 185 were greater then the
cumulative rainfall and the slopes of cumulative fluxes curves were steeper then that for
rainfall that indicated malfunctioning of PCAPS. For those reasons not all data were used,
but three wetting-drying periods were selected as shown in Fig. 3-14. Cumulative soil water
fluxes for those periods were used as three values of the key output.
3.4.2 Statistics to evaluate model abstractions
The F-test was used to estimate accuracy of flux simulation for each MA technique. The
test compares model error with the variation in the experimental data (Whitmore, 1991).
The sum of squares due to lack of fit was used to characterize the model error:
K
LOFIT = ∑ n j ( y j − x j ) 2
(3.1)
j =1
where nj is the number of replicates at each depth and time, xj is the simulated flux for each
depth and time, K is the total number of measurement points ( K = [number of depth] x
[number of time periods]), y j is the mean of the measurement at each depth and time:
73
60
Water Storage
Rainfall
40
40
38
20
36
0
34
-20
125
150
175
200
225
250
Cumulative rainfall (cm)
Water storage in soil profile (cm)
42
275
Julian day
Figure 3-14. Wetting-drying periods selected to compute the cumulative water flux as the
key output of the base and abstracted models.
74
n
1 j
∑ yij
n j i =1
where yij is the measured flux for each depth and time.
yj =
(3.2)
The sum of squares of the error was used to characterize variability of the measured data:
K
nj
SSE = ∑∑ ( y ij − y j ) 2
(3.3)
j =1 i =1
The F-statistics is computed as:
LOFIT
F=
(3.4)
SSE
The number of degrees of freedom in the LOFIT is K, and in the SSE is N-K, where N is the
total number of measurements including all replications.
Each abstracted model was tested to assess the propagation of the variability in parameters
through the model and the resultant variability in the key output and in simulated water
contents. The variability in parameters was inferred either form the inverse modeling
results or from the ensemble of pedotransfer functions as described below.. Probability
distribution functions of the root mean-square error in water contents and the probability
distribution functions of the key output were used to document and compare the error
propagation.
3.4.3 Available data and their uncertainty
The purpose of the data review is to establish model calibration dataset having reliable
statistics of all model inputs and outputs.
3.4.3.1 Averaging water contents
Inspection of time series of water content and pressure head revealed oscillatory behavior
that apparently had diurnal periodicity. The data analysis presented in the Appendix B
showed that that the oscillations of recorded capillary pressures and soil water contents
were related to the temperature on/near soil surface. That could be attributed to the
unknown response of the electronic equipment on the surface to the diurnal daily
oscillations. To eliminate temperature effect of measurement devices, daily time series of
water content and pressure head were averaged for each day of the observation period. The
temporal time scale of observations became equal to one day.
3.4.3.2 Temporal persistence in soil water contents
The water content data were incomplete due to malfunctioning of the measurement devices.
When removed from soil, all probes were still intact, and were subsequently used in other
laboratory or field experiments. The malfunctioning was apparently caused by the
connections and switch boxes in the automated measurement system that would stop
75
working properly resulting in missing data for the rest of the experiment for a given probe.
Averaging data from the remaining probes at a given depth could be an option if no
persistent difference had existed between TDR-measured water content in different
locations. However, the data inspection showed that, when all probes worked, some probes
at a given depth consistently showed water contents below average whereas others show
water contents above the average. Fig. 3-15 shows measurements made when all probes at
the 15-cm depth have worked. The graphs for the individual probes are clearly shifted
relative to each other along the water content axis. This demonstrates the temporal
persistence in water contents at different locations. Similar graphs were obtained for other
depths (data not shown).
Because of the temporal persistence, the malfunctioning of probes with consistently lower
than average values would lead to averaging only data from the remaining probes that had
water contents consistently higher than the true average. This averaging would lead to
overestimation of the average water content at the depth of interest. In contrary, the
malfunctioning of probes with consistently higher than average values would lead to the
underestimating the average.
The correction for the temporal persistence (Pachepsky et al., 2005) was applied to the data
as described in the Appendix C to estimate layer-averaged water contents. This correction
also led to the five-fold decrease in the uncertainty of the estimates of layer-averaged water
contents because, in case of persistence, the individual probe readings are not independent.
The depth-averaged water contents with the standard errors based on the persistence were
used in calibrations.
3.4.3.3 Surface evaporation
The calibration of the base model in (Jacques et al., 2002) was based on the assumption that
the surface evaporation was absent during the experiment. To evaluate the uncertainty
related to this assumption, we used the data from locations where the monitoring was
continuous at all observation depths. The cumulative change in water storage in soil profile
was computed for those locations. Results in Fig. 3-16 show that distinct periods of water
loss (“dry periods”) and water accumulation (“wet periods”) can be defined for all
locations. Inspection of capillary pressure and water content data during dry periods
showed that:
(a)
capillary pressure gradient at depths 65 cm and 85 cm was between zero and
one (Fig. 3-17); positive value of pressure gradient indicated downward
water flux from those layers with infiltration,
(b)
water content decreased and pressure head increased continuously at all
depths (Fig. 3-18)
(c)
water content increased linearly with depth in deep parts of soil profile (Fig.
3-19).
We used these data to estimate the matrix unsaturated hydraulic conductivity in the deep
part of the profile in the C2 soil horizon. We assumed that (a) no preferential flow occurred
76
0.40
a
0.35
-3
Volumetric water content (cm cm )
0.30
0.25
3
b
0.35
0.30
0.25
c
0.35
0.30
0.25
Daily rainfall (mm)
0.35
d
0.30
25
20
15
10
5
0
e
70
110
150
190
230
Julian day
Figure 3-15. Time series of TDR-measured water contents at the depth of 15 cm (a-d) and
precipitation (e) during the period when all probes worked at this depth. Probe numbers top
to bottom (a) 2, 3, 1, (b) 5, 6, 4, (c) 7, 8, 9, (d) 11, 12, 10; probe numbering is shown in
Figure 3-3.
77
Figure 3-16. Cumulative change in soil water storage in several locations along the trench. Location numbering is given in the Fig. 3-3.
78
Figure 3-17. Capillary pressure gradient during a drying period in the location 10
79
200
0.40
(a)
(b)
0.38
Pressure head (cm)
Water content
0.36
0.34
0.32
0.30
0.28
0.26
145
15
35
55
75
95
150
100
50
150
155
160
165
145
150
155
160
165
Time (day)
Time (day)
Figure 3-18. Daily average water content and pressure head time series at the location 10.
80
Figure 3-19. Water contents in soil profiles for several days of a drying period.
81
Figure 3-20. Estimated unsaturated hydraulic conductivity of soil matrix below 60 cm.
82
during the drying period, (b) the dependence of the unsaturated hydraulic conductivity on
water content could be assumed the same within the depth range from 65 to 85 cm, (c) the
linear interpolation was applicable to water contents. The procedure to estimate unsaturated
hydraulic conductivity is given in the Appendix D. It is based on the combining the mass
conservation equation and Darcy-Buckingham law and using the one-day time step. The
estimated dependence of the unsaturated hydraulic conductivity on volumetric water
content is shown in Fig. 3-20. This dependence was then used to compute the daily water
flux from the soil profile at the depth of 105 cm as shown in the Appendix D.
Being able to compute the matrix flux at the bottom of the soil profile and changes in water
storage of the soil profile, we could estimate evaporative losses during dry periods from the
water mass balance. Estimated daily evaporation for five dry periods is shown in Fig. 3-21.
The variability between dry periods is less than the variability within dry periods. The
Kruskal-Wallis rank sum test did not indicate (p-value = 0.924) difference in mean daily
evaporation values between dry periods, and ANOVA did not select the period as a
influential factor at 0.05 probability level. The probability distribution function of estimated
evaporation is given in Fig. 3-22. The values of daily evaporation were 0.29 mm, 0.86 mm
and 1.51 mm at probability levels 25%, 50% and 75%, respectively. About 10% of the
estimates are negative; this should be attributed to the error propagation, measurement
noise, and a large time step.
3.4.4 Additional information to ensure sound abstraction
The purpose of collecting the additional information is to ensure that the abstracted models
includes descriptions of all essential processes of flow and transport for given site.
Although this additional information may be of lower quality compared with the necessary
part of the database, it may show that some small-scale internal heterogeneities may have a
dominant effect on the flow.
3.4.4.1 Evidence of the macropore flow
The presence of the macropore flow was one of the hypothetical reasons why the base
model in HYDRUS-1D failed to simulate infiltration after intensive rainfall (Jacques et al.
2002). We analyzed the field water retention data and used the estimates of the surface
evaporation from the section 3.3.2.3 to see whether the macropore flow might be present.
Fig. 3-23 shows field water retention at the 15-cm depth for the drying period from day 140
to day 190. The drying retention curves except curves at locations 110 and 310 have a
horizontal segments at high saturations. At those segments, water content decreases and
capillary pressure remains approximately constant as soil dries. The capillary pressure
measured with the tensiometer reflects soil water potential in soil matrix. The decrease in
soil water contents reflects the loss of soil water in macropores outside soil matrix. This
loss occurs because both evaporation and redistribution. When the macropore water is lost,
evaporation begins to draw water from soil matrix. This affects the tensiometer readings,
and capillary pressure begins to rise simultaneously with the decrease of soil water
contents. The schematic of such soil drying is shown in Fig. 3-24. This type of drying
83
Daily evaporation (mm)
6
4
2
0
-2
-4
100
150
200
250
300
Day
Figure 3-21. Estimated daily evaporation for dry periods during the experiment.
99
Probability (%)
90
70
50
30
10
1
-0.2
0.0
0.2
0.4
Daily evaporation (cm)
Figure 3-22. Probability distribution function of estimated daily evaporation.
84
0.6
Pressure head (cm)
200
Location 10
150
100
50
0
0.25
150
0.30
0.35
0.25
0.30
0.35
Presure head (cm)
Location 160
0.31
0.32
0.33
0.34
0.35
0.36
Location 260
Location 210
100
50
0
0.26
200
0.28
0.30
0.32
0.34
0.32
0.36
0.40
Location 310
Pressure head (cm)
Location 110
Location 60
0.44
0.24
0.26
0.28
0.30
Location 410
Location 360
150
100
50
0
0.32
300
0.34
0.36
0.25
0.30
0.35
Location 460
0.28
0.32
0.36
0.40
Location 560
Location 510
Pressure head (cm)
250
200
150
100
50
0
0.32
0.34
0.36
0.38
0.40 0.34
0.36
0.38
0.40
3
0.32
0.34
0.36
0.38
0.40
3
Water content (cm /cm )
Figure 3-23. Relationships between pressure head and water content at the 15-cm depth
between 140 and 190 day from the beginning of the experiment. Location numbering is
given in the Fig. 3-3.
85
60
Water is released
from the soil matrix
(meso- or micropores),
tenziometer readings
change along with changes
in soil water content
Pressure head (cm)
50
40
30
Macroporosity
empties,
tensiometer readings
do not change
20
10
0
0.30
0.32
0.34
0.36
0.38
3
0.40
-3
Volumetric water content (cm cm )
Figure 3-24. Drying water retention curve of a soil with macroporosity.
86
0.42
Cumulative flux (cm)
Precipitation
Evaporation
Macropore flow
Matric flow
60
40
20
0
0
50
100
150
200
250
Time (day)
Figure 3-25. Cumulative water fluxes computed with constant evaporation of 0.86 mm day.
87
indicates the well-developed macroporosity. Therefore, soil at the site has necessary pore
structure to exhibit the fast water flow in macropores in close-to-saturation conditions if
case macropores have a good connectivity. Pore structure of the soil under study may serve
as a conduit for the preferential flow which is, by definition, the non-uniform downward
flow is substantially larger than the estimated matrix flow. It develops predominantly
during the intermediate and wet periods and generally follows the pattern of precipitation.
The estimates of water balance, hydraulic conductivity, and surface evaporation are not
necessarily accurate. Several strong, albeit plausible, assumptions have been made to obtain
them. However, those estimates show that the substantial macropore flow in the soil could
develop during intensive rainfall events. Macropores present small scale-heterogeneities in
pore space that can substantially affect flow and transport in variably saturated subsurface.
The abstracted models should include the ability to simulate macropore flow.
3.3.5 Recalibration of the base model
The Richards single continuum model was recalibrated using (a) water contents corrected
for daily oscillations and temporal persistence (Sections 3.4.3.1 and 3.4.3.2) and (b) the
estimated evaporation rate as boundary condition to see whether the inexplicable behavior
of the model will disappear. Soil profile was subdivided into five layers 0-25, 25-45, 45-65,
65-85, and 85-105 cm. Four parameters θr, θs, α, n of the van Genuchten equation (2-2) and
two parameters Ks, l of the Genuchten-Mualem equation (2-4) were estimated for each
layer fitting the HYDRUS-1D to the corrected averaged water content. The initial estimates
of the parameters were found using data on soil texture (Table 3-1) and the Rosetta
software included in the HYDRUS-1D model. The total 30 parameters were estimated
during the model calibration.
Calibration resulted in the relatively high accuracy (Fig. 3-26) of water contents. The
RMSE (root-mean-square error) was 0.0058 cm3cm-3, the determination coefficient of the
linear regression ‘observed vs. simulated’ water contents was R2 = 0.834. Calibrated
parameters along with their standard errors and linear tolerance intervals are shown in
Table 3-6. Parameter θr is found with the lowest accuracy because it affects the water
retention at high capillary pressures whereas the observed capillary pressures were very low
(Fig. 3-8). In contrary, values of θs are very reliable. Parameters α and n are defined reliably
only in the top layer. In all other layers, the tolerance intervals are very wide but the
estimates themselves are acceptable. As indicated by Hill (1998), this suggests that the data
in deep layers are insufficient for conclusive evaluation. And indeed, soil in top layer
experiences some drying (Fig. 3-26), and therefore simulations are sensitive to the retention
curve shape parameters α and n. In deeper layers, no substantial drying occurs, and the
retention curve shape parameters α and n cannot be found reliably because all experimental
points are located very close to the wet end. The same is true for the parameters of the
hydraulic conductivity curve Ksat and l.
Although the calibration results were satisfactory (Fig. 3-26), simulated soil water fluxes
were still substantially different from measured (Fig. 3-27). The F-statistic of 5.92
computed according Eq. (3-4) was above the critical value of 2.5. The Richards single
88
0.40
15 cm
0.38
0.36
0.34
0.32
0.30
35 cm
0.36
0.34
Water content (cm3.cm-3)
0.32
0.30
55 cm
0.37
0.36
0.35
0.34
75 cm
0.38
0.36
0.34
95 cm
0.38
0.36
Measured
HYDRUS-1D
0.34
0.32
100
150
200
250
300
350
400
450
Time (day)
Figure 3-26. Time series of the depth-averaged water contents; ○ – measured, ― simulated
with the calibrated Richards medium model.
89
Table 3-6. Calibrated van Genuchten (Eq. 2-2) and van Genuchten-Mualem (Eq.2-5)
parameters of the Richards medium model (HYDRUS-1D)
Variable
θr
3
θs
-3
3
α
-3
cm cm
cm cm
Estimate
Std. error
Lower 95%
Upper 95%
0.005
0.22
-0.427
0.437
0.363
0.001
0.361
0.365
Estimate
Std. error
Lower 95%
Upper 95%
0.086
1.409
-2.678
2.85
0.362
0.008
0.347
0.377
Estimate
Std. error
Lower 95%
Upper 95%
0.154
1.15
-2.102
2.41
0.367
0.004
0.359
0.375
Estimate
Std. error
Lower 95%
Upper 95%
0.108
1.443
-2.722
2.938
0.388
0.008
0.372
0.404
Estimate
Std. error
Lower 95%
Upper 95%
0.01
0.453
-0.878
0.898
0.386
0.004
0.378
0.394
n
-1
cm
0-25 cm
0.0011
0.0004
0.0003
0.0019
25-45 cm
0.00184
0.004
-0.0061
0.0098
45-65 cm
0.00123
0.00438
-0.0074
0.00982
65-85 cm
0.00133
0.00426
-0.007
0.00969
85-105 cm
0.00155
0.00161
-0.0016
0.0047
90
Ksat
l
-1
cm day
1.8
0.189
1.43
2.17
0.066
0.006
0.054
0.078
0.01
0.83
-1.63
1.64
1.267
1.136
-0.961
3.495
50
409.61
-753.34
853.34
0.5
36.4
-70.88
71.88
1.527
1.35
-1.12
4.174
0.207
0.3
-0.381
0.795
5
54.4
-101.7
111.7
1.492
1.441
-1.334
4.318
0.167
0.329
-0.478
0.812
0.5
24.87
-48.28
49.28
1.199
0.197
0.812
1.586
4.159
7.646
-10.836
19.154
18.4
31.5
-43.5
80.3
30
80
15 cm
55 cm
60
20
40
Simulated cumulative flux (cm)
10
20
0
0
0
10
20
30
30
0
20
40
60
80
Measured cumulative flux (cm)
105 cm
20
HYDRUS-1D
1:1 line
10
0
0
10
20
30
Measured cumulative flux (cm)
Figure 3-27. Soil water fluxes simulated with calibrated Richards medium model; error bars
show one standard deviation
91
continuum model predicted either water accumulation at the soil surface or simulated
substantial surface runoff during large storm events. As a result, simulated water fluxes
through soil were much less than measured with passive capillary lysimeters or estimated
from water balance. points are located very close to the wet end. The same is true for the
parameters of the hydraulic conductivity curve Ksat and l.
Although the calibration results were satisfactory (Fig. 3-26), simulated soil water fluxes
were still substantially different from measured (Fig. 3-27). The F-statistic of 5.92
computed according Eq. (3-4) was above the critical value of 2.5. The Richards single
continuum model predicted either water accumulation at the soil surface or simulated
substantial surface runoff during large storm events. As a result, simulated water fluxes
through soil were much less than measured with passive capillary lysimeters or estimated
from water balance.
Overall, the base model failed to simulate the key output when fitted to spatially and
temporally corrected data using corrected boundary conditions. Therefore, the need in
model abstraction persisted.
3.5 Selection and application of model abstraction techniques
3.5.1 Model abstraction design
The review of model abstraction techniques listed in Chapter 2 showed that the following
four categories of the techniques are applicable to this case study: (a) the abstraction with a
hierarchy of models, (b) the abstraction with scale change (aggregation); (b) the abstraction
of parameter determination with pedotransfer functions, and (d) the abstraction with
metamodeling. The operational scale was not changed, and no data were available to
perform discretization or scaling in parameter determination. Therefore, other categories of
abstraction were not used.
The design of the model abstraction process is shown in Fig. 3-28. Eight abstractions were
performed. The base model is shown as the model 0 in Fig. 3-28. The model 1 is the result
of model hierarchy abstraction, the model 2 is the result of the abstraction with scale
change (aggregation); the model 3 is the result of the abstraction of parameter
determination with pedotransfer functions, the model 4 is the result of the abstraction with
metamodeling. Another four models are built as a result of the sequential abstraction as
discussed below.
Model abstraction experiments were performed in the uncertainty context. Fifty simulations
were done for each abstracted model. The random input was generated according to the
statistical distributions of parameter values. Details of the Monte Carlo simulations are
given in following subsections.
3.5.2 Abstraction with a hierarchy of models: using water budget
92
The hierarchy in Fig. 2-2 suggests using water budget model as a simplification of the
single continuum, or Richards medium, model. The water budget model MWBUS (Model
of Water Budget of Unsaturated Soil) has been used in this work. Detailed description of
the model is in the Appendix E. The model is based on the model AQUA (Guber et al.,
1995). The model is conceptually similar to the model MEPAS (Whelan et al, 1996). The
difference is that MWBUS does not assume the steady-state flow and includes the depthdependent loss of water for evaporation and/or transport to upper layers as proposed by
Suleiman and Ritchie (2003). The model has the ability to simulate macropore flow as
required in Section 3.3.4.1. The MWBUS is used in models 1, 5, 7, and 8 shown in Fig. 328.
The MWBUS was calibrated on corrected averaged water content using rainfall and
estimated evaporation rate as upper boundary condition. The estimated parameters were
saturated water content θs and field capacity θfc for five layers, and parameters of the
evaporation equation (E4) a, b, and τ0. The values of saturated hydraulic conductivity were
assigned at 50% probability according to soil texture and porosity (Table 2-5) and were not
changed during optimization. Thus, the total number of estimated parameters was equal to
13. The version of Marquardt algorithm (van Genuchten, 1981) was used in the calibration.
Results of the MWBUS calibration are shown in Fig. 3-29 and 3-30. As compared with the
Richards medium model, the accuracy of the water content simulation with the calibrated
MWBUS was worse (compare Fig. 3-29 and Fig. 3-30), the RMSE was 0.00719 cm3cm-3
and the determination coefficient of the linear regression ‘observed vs. simulated’ water
contents was R2=0.740. However, the correspondence between MWBUS-simulated and
measured soil water fluxes was satisfactory (Fig. 3-28). The F-statistic value was below the
critical value (Fig. 3-31). Calibrated parameters of the MWBUS model and their statistics
are given in Table 3-7. The MWBUS parameters are much more reliable compared with the
parameters of the Richards model.
The abstraction from the Richards model to the water budget model not only provided a
reasonable estimation of the soil water fluxes as the key output, but also served as a “sanity
check” giving a clear indication that the Richards medium model was optimized in the
wrong domain of its parameter space.
3.5.3 Abstraction with the scale change: profile aggregation
The layered soil profile was replaced with homogeneous one in this abstraction (model 2 in
Fig. 3-28). The HYDRUS-1D was calibrated with averaged water content corrected as
described in Sections 3.3.2.1 and 3.3.2.2. Four parameters θr, θs, α, n of the van Genuchten
equation (2-2) and two parameters Ks, l of the Genuchten-Mualem equation (2-4) were
estimated for the whole soil profile fitting the HYDRUS-1D model. The total number of
estimated parameters was equal to 6. Results of the calibration are shown in Fig. 3-32.
Replacing the layered profile with homogeneous profile has more effect on errors in water
contents than on errors in flux computations (compare Fig 3-29 with 3-32, and Fig. 3-30
with Fig. 3-33). Water fluxes were not simulated correctly after this abstraction. The Fstatistic computed according Eq. 3-4 was above the critical value (Fig. 3-31).
93
Homogeneous
Layered
Scale change - aggregation
4
6
8
2
7
e
et
r
m
ra
a
p on
3
5
of nati
n i
tio rm
c
e
tra det
s
0
1
n
Ab
tio
a
r
lib
a
C
Abstraction of the model structure
Water
budget
Richards
medium
Figure 3-28. Design of the model abstraction application in this work. The Richards media
was simulated with the HYDRUS-1D software; the water budget model was MWBUS.
94
0.40
15 cm
0.38
0.36
0.34
0.32
0.30
35 cm
0.36
0.34
Water content (cm3.cm-3)
0.32
0.30
55 cm
0.37
0.36
0.35
0.34
75 cm
0.38
0.36
0.34
95 cm
0.38
0.36
Measured
MWBUS
0.34
0.32
100
150
200
250
300
350
400
450
Time (day)
Figure 3-29. Time series of the depth-averaged water contents; ○ – measured, ― simulated
with the calibrated MWBUS model.
95
30
80
15 cm
55 cm
60
20
40
Simulated cumulative flux (cm)
10
20
0
0
0
10
20
30
30
0
20
40
60
80
Measured cumulative flux (cm)
105 cm
20
HYDRUS-1D
MWBUS
10
0
0
10
20
30
Measured cumulative flux (cm)
Figure 3-30. Soil water fluxes simulated with calibrated HYDRUS-1D and the MWBUS
models; error bars show one standard deviation
96
Homogeneous
Layered
Scale change - aggregation
[6] 1.13¶
[8] 0.85¶
[7] 0.95¶
[2] 6.70
er
et
m
r
ra
fe
a
n
[3]
[5]
s
p
o
an ns
of nati
r
t
n
i
do ctio
tio rm
e
c
e
P fun
tra det
[0] 5.92
[1] 0.86¶
s
n
Ab
io
t
ra
lib
a
C
Abstraction of the model structure
1.14¶
Richards
medium
0.84¶
Water
budget
Figure 3-31. F-statistics of soil water flux simulations with different abstraction techniques
for depths of 15 and 55 cm.; ¶ - the evaluated F-value is less than the critical F-value at 95%
probability; model numbering in brackets is the same as in Fig. 3-28. The critical value of
F-statistics is 2.5.
97
Table 3-7. Calibrated parameters of the water budget model (MWBUS)
θs
θfc
Ksat
θr¶
cm3cm-3
cm3cm-3
cm day-1 cm3cm-3
Layered soil profile
0-25 cm
Estimate
0.432
0.353
133.9
0.09
Std. error
0
0.000
0
Lower 95%
0.432
0.353
133.9
Upper 95%
0.432
0.354
133.9
25-45 cm
Estimate
0.384
0.348
133.9
0.101
Std. error
0
0.000
0
Lower 95%
0.384
0.347
133.9
Upper 95%
0.384
0.349
133.9
45-65 cm
Estimate
0.388
0.360
133.9
0.108
Std. error
0
0.000
0
Lower 95%
0.388
0.359
133.9
Upper 95%
0.388
0.361
133.9
65-85 cm
Value
0.388
0.374
14.9
0.123
Std. error
0
0.000
0
Lower 95%
0.388
0.374
14.9
Upper 95%
0.388
0.375
14.9
85-105 cm
Estimate
0.388
0.370
14.9
0.109
Std. error
0
0.001
0
Lower 95%
0.388
0.368
14.9
Upper 95%
0.388
0.371
14.9
Homogeneous soil profile
Variable
k0
a
b
-1
day
day cm
Estimate
109.7
0.602
1.579
Std. error
3.2
0.023
0.012
Lower 95%
103.3
0.557
1.555
Upper 95%
116.0
0.646
1.603
¶
Was estimated from pedotransfer function Rosetta in HYDRUS-1D
Variable
98
3.5.4 Abstraction of parameter determination: using an ensemble of pedotransfer
functions
This abstraction corresponds to the model 3 in Fig. 3-28. Calibration was replaced
with using pedotransfer functions for both water retention and hydraulic conductivity.
Results of application of 21 pedotransfer functions given in the Appendix A and the Rosetta
software are shown in Fig. 3-34. None of the PTFs provide a good approximation of the
field water retention, although the field variability makes questionable any attempt of PTFbased predictions, just because no data on PTF inputs collocated with the measurement
points have been available. Nevertheless, the ensemble of pedotransfer functions creates an
envelope that encompasses the range of field water retention data. Similar results were
obtained at other depths (data not shown).
The Monte Carlo simulations were performed with HYDRUS-1D model using each
pedotransfer function from the Appendix 1 and the Rosetta pedotransfer function to
estimate water retention. Table 3-8 indicates whether Brooks-Corey equation (2-1) or van
Genuchten equation (2-2) was used in simulations with a specific pedotransfer function.
The median values of lognormal distributions of Ksat shown in Table 2-5 and in Fig. 2-10
were used from loam and silt loam texture classes at depths ranges 0 to 65 cm and 65 to
105 cm, respectively. Results of water content simulations are shown in Fig. 3-35. The
accuracy of those simulations in terms of water contents was low. There has obviously been
a mismatch between the simulated initial distributions of pressure heads and the actual
ones, and the first period of simulations was characterized by the substantial loss of water
through the bottom of the profile. However, the accuracy of soil water fluxes estimates in
simulations with pedotransfer functions was satisfactory (Fig. 3-36). The F-statistic value
was below the critical value (Fig. 3-31). These results show that the Richards media model
is, in principle, capable of correct prediction of soil water fluxes in this case study provided
realistic parameter estimates are used.
3.5.5 Sequential model abstractions
Models 5, 6, and 7 in Fig. 3-28 represent application of two model abstraction techniques.
The purpose of such sequential application is to find out whether additional gains can be
obtained without the loss in accuracy.
Model 5 in Fig. 3-28 represents using pedotransfer functions with soil water budget model.
The Monte Carlo simulations were performed with the MWBUS model using each
pedotransfer function from the Appendix 1 and the Rosetta pedotransfer function to
estimate water retention. Water retention was estimated using average soil texture for depth
15 and 35 cm for topsoil, and 55, 75, 95 cm for subsoil. The Monte Carlo simulations
required a decision on how to estimate field capacity from such data. Although there is no
99
0.40
15 cm
0.38
0.36
0.34
0.32
0.30
35 cm
0.36
0.34
Water content (cm3.cm-3)
0.32
0.30
55 cm
0.37
0.36
0.35
0.34
75 cm
0.38
0.36
0.34
95 cm
0.38
0.36
Measured
HYDRUS-1D
MWBUS
0.34
0.32
100
150
200
250
300
350
400
450
Time (day)
Figure 3-32. Time series of transect-average water contents: symbol – measured, red lines –
calibrated the HYDRUS-1D model, blue lines - calibrated the WMBUS model for
homogeneous soil profile.
100
80
30
15 cm
55 cm
60
20
40
Simulated cumulative flux (cm)
10
20
0
0
0
10
20
30
30
0
20
40
60
80
Measured cumulative flux (cm)
105 cm
20
HYDRUS-1D
MWBUS
10
0
0
10
20
30
Measured cumulative flux (cm)
Figure 3-33. Cumulative water fluxes simulated with calibrated HYDRUS-1D and
MWBUS models for homogeneous soil profile; error bars show one standard deviation.
101
Capillary pressure (cm)
300
200
100
0
0.2
0.3
0.4
Water content (cm3cm-3)
Figure 3-34. Comparison of field water retention (symbols) at the depth of 15 cm with
water retention estimated using 21 pedotransfer functions listed in the Appendix A and the
Rosetta software.
102
Table 3-8. PTF used to run HYDRUS 1D with Brooks-Corey (BC) and Van Genuchten
(VG) water retention equations
# Authors
Clay Silt
Sand BD¶ OM§ topsoil
1 Campbell & Shiozawa (1992)
BC
+
+
+
2 Rawls & Brakensiek (1989)
BC
+
+
+
3 Saxton et al. (1986)
BC
+
+
4 Oosterveldt & Chang (1980)
BC
+
+
5 Williams et al. (1992)
BC
+
+
+
6 Williams et al. (1992)
BC
+
+
+
7 Baumer (1992)
VG
+
+
+
+
8 Bruand et al. (1994)
VG
+
9 Canarache (1993)
VG
+
+
10 Gupta & Larson (1979)
VG
+
+
+
+
+
11 Hall et al. (1977)
VG
+
+
+
+
12 Mayr & Jarvis (1999)
VG
+
+
+
+
+
13 Pachepsky et al. (1982)
VG
+
+
+
14 Petersen et al. (1968)
VG
+
15 Rajkai & Varallyay (1992)
VG
+
+
+
+
16 Rawls et al. (1982)
VG
+
+
+
+
17 Rawls et al. (1992)
VG
+
+
+
+
+
18 Verekeen et al. (1989)
VG
+
+
+
+
19 Tomasella & Hodnett (1998)
VG
+
+
20 Wosten et al. (1999)
VG
+
+
+
+
21 Wosten et al. (1999)
VG
+
+
+
+
+
+
22 Rosetta
VG
+
+
+
+
¶
Average bulk density values were taken from Mallants et al. (1996) as 1.426, 1.544, and
1.526 g cm-3 in the depth ranges 0 to 25 cm, 25 to 55 cm and 55 to 105 cm, respectively
§
Organic matter content was set at 2% in the 0-25 cm topsoil, and 0.5% in subsoil.
103
0.40
15 cm
0.35
0.30
0.25
0.20
0.15
35 cm
0.35
0.30
Water content (cm3.cm-3)
0.25
0.20
55 cm
0.35
0.30
0.25
0.20
75 cm
0.35
0.30
0.25
95 cm
0.35
0.30
0.25
100
150
200
250
300
350
400
450
Time (day)
Figure 3-35. Time series of transect-average water contents: symbol – measured, red lines simulated with the HYDRUS-1D, blue lines - simulated with the WMBUS model; water
retention data and median hydraulic conductivity estimated with pedotransfer functions; the
solid line corresponds to the median value of simulated water contents for each day, dashed
lines correspond to the simulated daily values at 25% and 75% probability levels.
104
80
30
15 cm
55 cm
60
20
40
Simulated cumulative flux (cm)
10
20
0
0
0
10
20
30
30
0
20
40
60
80
Measured cumulative flux (cm)
105 cm
20
HYDRUS-1D
MWBUS
10
0
0
10
20
30
Measured cumulative flux (cm)
Figure 3-36. Soil water fluxes simulated with HYDRUS-1D and MWBUS models; water
retention data and median hydraulic conductivity estimated with pedotransfer functions for
layered soil profile; error bars show one standard deviation.
105
unique capillary pressure value to estimate field capacity from ‘water content-capillary
pressure’ dependencies (see section 2.2 and Fig. 2-6), two often-suggested values of soil
water potential for such estimate are 100 cm and 330 cm. We used the laboratory water
retention data to select the capillary suction corresponding to the field capacity. The
distributions of water content at both capillary pressures were constructed by random
sampling the distributions of van Genuchten parameters shown in Fig. 3-12, and the soil
water was simulated with both distributions. The root-mean-square errors with water
contents at capillary pressure of 100 cm as field capacity values were about two times less
than the errors with water contents at capillary pressure of 330 cm (data not shown).
Therefore, to perform the MWBUS model with parameters estimated using PTF (Table 38), the values of θfc were evaluated at pressure head of 100 cm. Parameters of the
evaporation model for the MWBUS were taken from the calibration as described in 3.5.2
Simulations were run for each function with the median values of lognormal distributions
of Ksat shown in Table 2-5 and in Fig. 2-10 that were used from loam and silt loam texture
classes at depths ranges 0 to 65 cm and 65 to 105 cm, respectively. Results of water content
simulations are shown in Fig. 3-35. The accuracy of those simulations was low. There has
obviously been a mismatch between water field capacity estimated with PTF and the actual
ones. The first period of simulations was characterized by the substantial loss of water
through the bottom of the profile. However, the accuracy of soil water fluxes estimates in
simulations with pedotransfer functions was satisfactory (Fig. 3-36).
Model 6 in Fig. 3-28 represents using pedotransfer functions after the aggregation of the
soil profile is performed. Soil texture was different for topsoil and subsoil in the soil profile
(Fig. 3-3). Pedotransfer functions require the soil textural composition as the input.
Therefore two cases were considered separately: homogeneous soil profile with topsoil
hydraulic properties, homogeneous soil profile with subsoil hydraulic properties.
Results of simulations are shown in Fig. 3-37 and 3-38. Simulated water content and water
fluxes were similar to those, obtained for layered soil profile (Fig. 3-35 and 3-36). The
Richards media model underestimated water content at all depths. The accuracy of flux
predictions was acceptable. The F-statistic computed for water fluxes was below the critical
value (Fig. 3-31). Replacing topsoil hydraulic parameters with subsoil parameters did not
result in noticeable changes in water content and flux values for both models (data not
shown). That might be due to variability in water retention caused by different PTF is
greater than that caused by changes in soil texture for particular soil profile.
Overall, the model abstraction with scale change (aggregation) resulted in deterioration of
the accuracy of soil water content simulations with both HYDRUS-1D and MWBUS
models. Soil water fluxes, however, were simulated well on both abstraction steps.
Model 7 in Fig. 3-28 represents the sequential use of hierarchy of models and aggregation.
The MWBUS has the single homogeneous layer and is calibrated for this case. Calibrated
parameter values and their statistics are shown in Table 3-7. The tolerance intervals are
narrow, parameter values are robust. Two parameters - saturated water content θs and field
capacity θfc, - and parameters of the evaporation equation (E4) a, b, and τ0 were estimated
fitting the MWBUS model. The values of saturated hydraulic conductivity were assigned at
106
0.40
15 cm
0.35
0.30
0.25
0.20
35 cm
0.35
0.30
Water content (cm3.cm-3)
0.25
0.20
55 cm
0.35
0.30
0.25
0.20
75 cm
0.35
0.30
0.25
0.20
95 cm
0.35
0.30
0.25
0.20
100
150
200
250
300
350
400
450
Time (day)
Figure 3-37. Time series of transect-average water contents: symbol – measured, red lines simulated with the HYDRUS-1D, blue lines - simulated with the WMBUS model for
homogeneous soil profile; water retention data and median hydraulic conductivity
estimated with pedotransfer functions for topsoil; the solid line corresponds to the median
value of simulated water contents for each day, dashed lines correspond to the simulated
daily values at 25% and 75% probability levels.
107
80
30
15 cm
55 cm
60
20
40
Simulated cumulative flux (cm)
10
20
0
0
0
10
20
30
30
0
20
40
60
80
Measured cumulative flux (cm)
105 cm
20
MWBUS
HYDRUS-1D
10
0
0
10
20
30
Measured cumulative flux (cm)
Figure 3-38. Soil water fluxes simulated with HYDRUS-1D and the MWBUS model; water
retention data and median hydraulic conductivity estimated with pedotransfer functions for
topsoil profile; error bars show one standard deviation.
108
80
30
15 cm
55 cm
60
20
40
Simulated cumulative flux (cm)
10
20
0
0
0
10
20
30
30
0
20
40
60
80
Measured cumulative flux (cm)
105 cm
20
HYDRUS-1D
MWBUS
10
0
0
10
20
30
Measured cumulative flux (cm)
Figure 3-39. Soil water fluxes simulated with calibrated HYDRUS-1D and the MWBUS
model for the homogeneous soil profile; error bars show one standard deviation.
109
50% probability according to soil texture and porosity (Table 2-4) and were not changed
during the MWBUS optimization. Thus, the total number of estimated parameters was
equal to 5.
Results of the calibration are shown in Fig. 3-32 and 3-33. Replacing the layered profile
with homogeneous profile has more effect on errors in water contents than on errors in flux
computations (see Fig. 3-29 and 3-30). The MWBUS model simulated water content
qualitatively less adequately compared to the MWBUS. The values of water content
simulated with the MWBUS did not changed in time at depths 55, 75 and 95 cm and were
equal to field capacity. However, the MWBUS model simulated water fluxes more
accurately compared to the HYDRUS-1D model when calibrated for homogeneous soil
profile (Fig. 3-39).
Model 8 in Fig. 3-28 represents the sequential application of three abstraction steps:
hierarchy of models, aggregation and pedotransfer functions. This ultimate simplification
leads to further deterioration of the accuracy in water content estimates (Fig. 3-37).
However, soil water fluxes are estimated reasonable well (Fig. 3-38).
3.5.6 Abstraction with metamodeling
This type of MA should be used when a large number of model runs may be required for
the performance assessment (see Section 2.2.4). We developed an example to show the
feasibility of applying an artificial neural network (ANN) to mimic the model behavior in
predicting soil water fluxes with meteorological inputs defined at daily, weekly and ten-day
time scales. The design of this work is shown in Fig. 3-40. It includes generating probable
weather patterns, running the models in the Monte Carlo simulation mode, dividing the data
into development and testing subsets, training the neural net to mimic the model output on
the development data subset, and testing the net on the testing data subset.
We used the feed-forward ANN as described by Pachepsky et al. (1996). Climatic
parameters of Albany, NY were used to generate 300 daily precipitation and evaporation
sequences for the April-September period. The sequences were randomly divided into 150
sequences for training and 150 sequences for testing. Feedforward networks were
developed using the MATLAB Neural Network toolbox. The minimum number of the
hidden neurons was found experimentally to provide the best accuracy with testing dataset.
ANN inputs were daily average precipitation and evaporation, or weekly average
precipitation and evaporation. Accordingly, ANN outputs were daily soil water fluxes or
weekly soil water fluxes.
Results are shown in Fig. 3-41 – 3-44. ANN related inputs and outputs of the models well.
The determinations coefficients of regressions ‘ANN monthly vs. simulated monthly’
ranged from 0.899 to 0.944 when daily inputs and outputs were used with the MWBUS
model (Fig. 3-41). There was a marked increase in accuracy of monthly fluxes predictions
when weekly input and output were used; the determination coefficients were between
0.959 and 0.971 (Fig. 3-42). There was an opposite trend with the HYDRUS-1D output.
The accuracy was higher with the daily input and outputs were used; values of R2 were in
the range between 0.990 and 0.994 (Fig. 3-43). With weekly inputs and outputs, the
110
Figure 3-40. Schematics of the neural network application to abstract simulation models.
111
20
Month 1
Month 2
Month 3
Month 4
Month 5
Month 6
15
10
5
Simulated monthly fux (cm)
0
20
15
10
5
0
20
15
10
5
0
0
5
10
0
15
5
10
15
20
ANN monthly fluxes (cm)
Figure 3-41.Estimated with the artificial neural network and simulated monthly soil water
fluxes at the depth of 105 cm. Model – MWBUS, artificial neural network input – daily
average flux, ANN output – daily average flux, the total number of hidden neurons – 20.
112
20
Month 1
Month 2
Month 3
Month 4
Month 5
Month 6
15
10
5
Simulated monthly flux (cm)
0
20
15
10
5
0
20
15
10
5
0
0
5
10
0
15
5
10
15
20
ANN monthly flux (cm)
Figure 3-42.Estimated with the artificial neural network and simulated monthly soil water
fluxes at the depth of 105 cm. Model – MWBUS, artificial neural network input – daily
average flux, ANN output – daily average flux, the total number of hidden neurons – 12.
113
20
Month 1
Month 2
Month 3
Month 4
Month 5
Month 6
15
10
5
Simulated monthly flux (cm)
0
20
15
10
5
0
20
15
10
5
0
0
5
10
0
15
5
10
15
20
ANN monthly flux (cm)
Figure 3-43. Estimated with the artificial neural network and simulated monthly soil water
fluxes at the depth of 105 cm. Model – HYDRUS-1D, artificial neural network input –
daily average flux, artificial neural network output – daily average flux, the total number of
hidden neurons – 20.
114
20
Month 1
Month 2
Month 3
Month 4
Month 5
Month 6
15
10
5
Simulated monthly flux (cm)
0
20
15
10
5
0
20
15
10
5
0
0
5
10
15
0
5
10
15
20
ANN monthly flux (cm)
Figure 3-44. Estimated with the neural network and simulated monthly soil water fluxes at
the depth of 105 cm. Model – HYDRUS-1D, ANN input – weekly average flux, ANN
output – weekly average flux, number of hidden neurons – 12.
115
accuracy slightly decreased, and more scatter could be seen in the 1:1 diagrams of
simulated versus ANN fluxes (Fig. 3-44). Values of values of R2 were in the range between
0.976 and 0.985 in this case.
Overall, the feedforward neural networks demonstrated an excellent ability to mimic the
simulations of soil water fluxes in bare soil under humid climatic conditions.
3.5.7 Pedotransfer-estimated vs. laboratory-measured water retention
Application of pedotransfer functions is often suggested only when measurements of
hydraulic properties are not available or are not feasible. Field measurements of soil water
retention were available in this case study. Additional simulations were made to estimate
the loss of accuracy caused by supplanting PTF-estimated water retention instead of the
measured one.
The Monte Carlo simulation was performed for the HYDRUS-1D model with van
Genuchten parameters shown in Fig. 3-12 and parameters estimated with PTF (Table 3-8).
No correlations between parameters shown in the Table 3-5 were used at this time. The
empirical distribution functions were sampled randomly 50 times. The median values of
lognormal distributions of Ksat (Table 2-5) were used from loam and silt loam texture
classes at depths ranges 0 to 65 cm and 65 to 105 cm, respectively.
The comparison of the statistical distributions of the root-mean-square errors obtained in
simulations with laboratory data and with pedotransfer functions showed that the accuracy
of simulations with PTF was better than with laboratory data (Fig. 3-45). In terms of
uncertainty no difference between simulation with laboratory and PTF were obtained, and
slight less uncertainty or more narrow range of RMSE for water content simulation was
obtained in case of layered soil profile.
The Monte-Carlo simulations with the MWBUS model were performed using water
contents at capillary pressure of 100 cm as the field capacity values. The probability
distributions of the MWBUS model root-mean-square errors from simulations with
laboratory data and with PTF functions showed that the accuracy of simulations with
laboratory data was better than with PTF (Fig. 3-45). Water content simulation results were
also less uncertain with the laboratory measured water retention than with use PTF.
The probability distributions were similar for layered and homogeneous soil profile with
PTF estimated water retention. For laboratory measured water retention more accurate but
less uncertain results were obtained for the homogeneous than for layered soil profile.
116
99
HYDRUS-1D
Probability (%)
90
70
50
30
10
1
0.00
0.05
0.10
0.15
0.20
99
MWBUS
Probability (%)
90
70
50
Water retention, Soil profile
30
Lab. measured, layered soil
Lab. measured, single layer
PTF estimated, layered soil
PTF estimated, single layer
10
1
0.00
0.05
0.10
0.15
3
0.20
-3
RMSE (cm cm )
Figure 3-45. Probability distribution of the root-mean square error (RMSE) of water
content from Monte Carlo simulations for the HYDRUS-1D and the MWBUS models with
two sources of soil hydraulic properties and two representations of soil profile.
117
3.6 Summary and conclusion
The case study elucidated the model abstraction efficiency in modeling water flow in
variably saturated subsurface. The intensive monitoring per se did not warrant the adequate
performance of the base model which was the Richards single continuum model. The need
in model abstraction was justified by the fact that the calibrated base model predicted
substantial runoff whereas no runoff was observed at the site during the monitoring period.
The base model produced inexplicable results in terms of the key output.
The review of modeling context helped to improve knowledge about boundary fluxes that
have not been measured. In particular, the review of the modeling context revealed that (a)
soil water contents across the trench demonstrated strong temporal persistence, and (b) a
substantial preferential flow should occur in soil. The persistence was used to estimate soil
bottom flux from water content measurements, and the preferential flow estimates were
used to assess evaporation from soil surface.
Four classes of applicable model abstraction techniques were selected. Model structure
abstraction was achieved using the hierarchy of model, aggregation, and metamodeling.
The hierarchy of models included the Richards single continuum medium and a simple soil
water budget model. Aggregation was done by replacing the layered soil profile with the
homogeneous profile. Metamodeling was done with the artificial neural network that used
daily and weekly precipitation as input to estimate monthly cumulative soil water fluxes.
An ensemble of 22 pedotransfer functions was used in model abstraction of parameter
determination. HYDRUS1-D, MWBUS, and MATLAB software packages were used to
calibrate the abstracted models and to run Monte Carlo simulations to evaluate the
uncertainty in calibrated model outputs.
The abstraction with the hierarchy of models was useful. The simple soil water budget
model was less accurate in predictions of soil water content compared with the Richards
model. However, unlike the more complex Richards model, this simple soil water budget
model correctly predicted the absence of runoff and measured cumulative soil water fluxes.
The prediction of runoff was an artifact of the Richards model calibration in absence of
measured boundary fluxes. This abstracted model appeared to be instrumental in both
explaining behavior of the complex model and in predicting the key output, i.e. soil water
fluxes.
The abstraction with aggregation was not useful in this case. The Richards model was less
accurate with respect to soil water contents and continued to generate large simulated
runoff when a homogeneous soil layer was introduced.
The abstraction of parameter determination with pedotransfer functions was useful. It
showed that the Richards model with parameters in correct ranges is able to correctly
simulate soil water fluxes. The ensemble of pedotransfer functions represented field water
retention better than data from laboratory soil water retention measurements. Soil water
content predictions with the Richards model were more accurate with the ensemble of
pedotransfer functions than with laboratory data.
118
The neural network metamodel was extremely accurate in estimating cumulative soil water
fluxes. It was also five orders of magnitude faster than the numerical algorithm coded in
HYDRUS-1D. However, the HYDRUS-1D runtime was adequate for the Monte Carlo
simulations to evaluate model abstraction in the uncertainty context.
In general, the model abstraction process was successful in this case study. Model
abstraction explained the strange behavior of the complex model, and provided the correct
description of the system behavior and plausible parameter ranges.
119
4 REFERENCES
Abbaspour, K.C., van Genuchten, M.T., Schulin, R., and E. Schläppi, “A Sequential
Uncertainty Domain Inverse Procedure for Estimating Subsurface Flow and
Transport Parameters,” Water Resources Research, 33(8): 1879–1892, 1997.
Addanki, S., Cremonini, R. and J. Penberthy, “Graphs of Models,” Artificial Intelligence,
51: 145-177, 1991.
Ahuja, L. R., J. W. Naney, P. E. Green, and D.R. Nielsen, “Macroporosity to Characterize
Spatial Variability of Hydraulic Conductivity and Effects of Land Management,”
Soil Science Society of America Journal, 48: 404-411, 1984.
Ahuja, L.R., D.K. Cassel, R.R. Bruce, and B.B. Barnes, “Evaluation of Spatial Distribution
of Hydraulic Properties Using Effective Porosity Data,” Soil Science, 148: 404-411,
1989.
Alexander, L., and R. W. Skaggs, “Predicting Unsaturated Hydraulic Conductivity From
Soil Texture,” Journal of Irrigation and Drainage Engineering, 113: 184-197, 1987.
Altman, S. J., Arnold, B. W., Barnard, R. W., Barr, G. E., Ho, C. K., McKenna, S. A., and
R. R. Eaton, “Flow Calculation for Yucca Mountain Groundwater Travel Time,”
(GWTT-95). SAND96-0819, Sandia National Laboratories, Albuquerque, NM, 212
pp., 1996.
Amemiya, M., “The Influence of Aggregate Size on Soil Moisture Content Capillary
Conductivity Relations,” Soil Science Society of America Proceedings, 29: 744-748,
1965.
Baumer, O. M. “Predicting Unsaturated Hydraulic Parameters,” pp.341-354. in M.Th. van
Genuchten, F. J. Leij, and L J. Lund (ed.) Proc. Int. Workshop on Indirect Methods
for Estimating the Hydraulic Properties of Unsaturated Soils. University of
California, Riverside, CA, 1992.
Benson, D.A., Wheatcraft S.W., and M.M. Meerschaert, “Application of a Fractional
Advection-Dispersion Equation,” Water Resources Research, 36: 1403-1412, 2000.
Berkowitz, B., Kosakowski, G., Margolin, G., and H. Scher, “Application of Continuous
Time Random Walk Theory to Tracer Test Measurements in Fractured And
Heterogeneous Porous Media,” Ground Water, 39: 593-604, 2001.
Beven, K. “Towards a Coherent Philosophy for Modeling the Environment,” Proceedings
of the Royal Society of London. Series A, Mathematical and Physical Sciences,
458(2026): 2465-2484, 2002
Bigelow, J. H,. and Davis, P. K.,”Implications for model validation of multiresolution,
multiperspective modeling (MRMPM) and exploratory analysis. MR-1750,” RAND,
2138, Santa Monica, CA, 2003
Bird, N. R. A, and E. M. A. Perrier, “The Water Retention Curve for a Model of Soil
Structure With Pore And Solid Fractal Distribution,” European Journal of Soil
Science,” 55: 55-67, 2000.
Bloemen, G. W., “Calculation of Hydraulic Conductivities of Soils From Texture and
Organic Matter Content,” Z. Pflanzenernahr. Bodenkd., 143: 581-605, 1980.
Blöschl, G., and Sivapalan, M., "Scale issues in hydrological modeling", Hydrological
processes, 9: 251-290, 1995.
121
Boyle, D. P., Gupta, H. V., Soroshian, J., Koren, V., Zhang, Z., M. Smith, “Toward
Improved Streamflow Forecasts: Value of Semidistributed Modeling,” Water
Resources Research, 37: 2749-2759, 2001.
Brooks, R.H., and A.T.Corey, “Hydraulic Properties of Porous Media,” Hydrology Pap. 3,
Colorado State University, Ft. Collins, 27 pp.,1964.
Bruand, A., Baize, D., and Hardy, M. 1994. Predicting of water retention properties of
clayey soil using a single soil characteristic. Soil Use and Management, 10: 99-103.
Burton, R., “Metamodeling: A State of The Art Review,” pp. 237-244 in: Proceedings of
the 1994 Winter Simulation Conference (J. D. Tew, S. Manivannan, D. A.
Sadowski, and A. F. Seila - Eds). Institute of Electrical and Electronics Engineering,
San Francisco, California. 1994.
Campbell, G.S., and S. Shiozawa, “Prediction of Hydraulic Properties of Soils Using
Particle Size Distribution and Bulk Density Data,” pp.317-328 in M. Th. van
Genuchten, F. J. Leij, and L J. Lund (ed.) Proceedings of the International
Workshop on Indirect Methods for Estimating the Hydraulic Properties of
Unsaturated Soils. University of California, Riverside, CA, 1992.
Canarache, A., “Physical-Technological Maps - A Possible Product of Soil Survey for
Direct Use in Agriculture,” Soil Technology, 6: 3-16, 1993.
Cassel, D. K., Wendroth, O., and D.R. Nielsen, “Assessing Spatial Variability in an
Agricultural Experiment Station Field: Opportunities Arising From Spatial
Dependence,” Agronomy Journal, 92: 706-714, 2000.
Castleton, K., Fine, S., Banta, N., Hill, M., Markstrom, S., Leavesley, G., and Babendreier,
J., “An Overview of the Uncertainty Analysis, Sensitivity Analysis, and Parameter
Estimation (UA/SA/PE) API and How To Implement It”, Pp. 171-172. In:
Nicholson, T. J., Babendreier, J. E., Meyer, P. D., Mohanty, S., Hicks, B. B.,
Leavesley, G. H. (Eds.) Proceedings of the International Workshop on Uncertainty,
Sensitivity, and Parameter Estimation for Multimedia Environmental Modeling
August 19–21, 2003. NUREG/CP-0187, U.S. Nuclear Regulatory Commission,
Washington, DC, 2002.
Caughlin, D. and A. F. Sisti, “Summary of Model Abstraction Techniques,” Proc. SPIE Int.
Soc. Opt. Eng. 3083: 2-13, 1997.
Christakos, G. Modern Spatiotemporal Geostatistics, Oxford University Press, Oxford,
New York, 212 pp., 2000.
Clancy, D., and B. Kuipers, “Model Decomposition and Simulation,” Proceedings of the
Eighth International Workshop on Qualitative Reasoning about Physical Systems,
Nara, Japan, pp. 45-54, 1994.
Clark, L. A. and D. Pregibon, “Tree-based Models,” pp. 377-419 in S.J.M. Chambers and
T. J. Hastie (eds.) Statistical Models, Wadsworth, 1991.
Clifton P. M., and S. P. Neuman, “Effects of Kriging and Inverse Modeling on Conditional
Simulation of The Avra Valley Aquifer in Southern Arizona,” Water Resource
Research, 18(4): 1215-1234, 1982.
Comegna, V. and A. Basile, “Temporal Persistence of Spatial Patterns of Soil Water
Storage in a Cultivated Vesubian Soil,” Geoderma 62, 299-310, 1994.
Cushman, J.H., “An Introduction to Fluids in Hierarchical Porous Media,” pp. 1-6 in J.H.
Cushman, ed., Dynamics of Fluids in Hierarchical Porous Media, Academic, NY,
1990
122
Davis, J. A., Yabusaki, S. B., Steefel, C. I., Zachara, J. M., Curtis, G. P., Redden, G. D.,
Criscenti, L. J. and B. D.Honeyman, “Assessing Conceptual Models for Subsurface
Reactive Transport of Inorganic Contaminants,” EOS, 85(44): 449-455, 2004.
Davis, P. K., and Bigelow, J. H,.”Motivated metamodels: synthesis of cause-effect
reasoning and statistical metamodeling. MR-1570,” RAND, 2138, Santa Monica,
CA, 2003.
Destouni, G. “Applicability of The Steady State Flow Assumption for Solute Advection in
Field Soils,” Water Resources Research, 27: 2129-2140, 1991.
Diwekar, U. M. “How Simple Can It Be? - A Look at The Models for Batch Distillation,”
Computers & Chemical Engineering 18: S451—S457, 1994.
Englert, A., Hardelauf, H., Vanderborght, J., and H. Vereecken, “Numerical Modelling of
Flow and Transport on Massively Parallel Computers,” NIC Symposium 2004,
Proceedings, D. Wolf, G. M¨unster, and M. Kremer (Editors), John von Neumann
Institute for Computing, Jülich, NIC Series, Vol. 20, ISBN 3-00-012372-5, pp. 409418, 2004.
European Soil Bureau, 1998. Georeferenced soil database for Europe. European Soil
Bureau and Joint Research Centre, EC. Ispra, Italy.
Ewing, R.E., Helmig, R., Finsterle, S., and R. Hinkelmann, “Modeling of Flow and
Transport Processes in the Subsurface,” pp. 477-480 in Groundwater Updates, K.
Soto and Y. Iwasa, eds., Tokyo, Springer-Verlag, 2000.
Falkenhainer, B. and K. Forbus, K., “Compositional modeling: finding the right model for
the job, “ Artificial Intelligence, 51 :95-143, 1991
FAO, 1975. Soil Map of the World at 1:5,000,000. Volume I. Europe. UNESCO, Paris,
France, 62 p.
Fayer, M.J. and C.S. Simmons, “Modified Soil Water Retention Functions for all Matric
Suctions,” Water Resources Research, 31: 1233-1238, 1995.
Ferreyra, R. A., Apeztegua, H. P., Sereno, R., and J. W. Jones, “Reduction of Soil Water
Spatial Sampling Density Using Scaled Semivariograms and Simulated Annealing,”
Geoderma, 110: 265-289, 2002.
Field, M. S., and P. F. Pinsky. "A two-region nonequilibrium model for solute transport in
solution conduits in karstic aquifers", Journal of Contaminant Hydrology, 44: 329351, 2000
Fine, S., Catleton, K, Banta, N., Markstrom, S., Leavesley, G., Babendreier, J.,
“Calibration, Optimization, and Sensitivity and Uncertainty Algorithms Application
Programming Interface (COSU-API), Pp. A1-A54. In: Nicholson, T. J.,
Babendreier, J. E., Meyer, P. D., Mohanty, S., Hicks, B. B., Leavesley, G. H. (Eds.)
Proceedings of the International Workshop on Uncertainty, Sensitivity, and
Parameter Estimation for Multimedia Environmental Modeling August 19–21, 2003.
NUREG/CP-0187, U.S. Nuclear Regulatory Commission, Washington, DC, 2002.
Fishwick, P. A. “The Role of Process Abstraction in Simulation,” IEEE Transactions on
Systems, Man and Cybernetics, 18: 18–39, 1988.
Fishwick, P. A. and P. A. Luker, (Eds.), “Qualitative Simulation Modeling and Analysis,”
Springer Verlag, 1991, 330 pp., 1991.
Fishwick, P. A., “Simulation Model Design and Execution”. Englewood Cliffs: PrenticeHall, 1995
123
Flury, M. and H. Flühler, “Modeling Solute Leaching in Soils by Diffusion-Limited
Aggregation: Basic Concepts and Application to Conservative Solutes,” Water
Resources Research, 31: 2443-2452, 1995.
Frantz, F. K. “A Taxonomy of Model Abstraction Techniques,” 2002.
http://www.rl.af.mil/tech/ papers/ModSim/ModAb.html
Frantz, K. F., “A taxonomy of model abstraction techniques,” In: Alexopoulos, C., Kang,
K., Lilegdon, W. R., and Goldsman, D. (Eds.) Proceedings of the 1995 Winter
Simulation Conference, pp.1413-142, 1995.
Franzmeier, D.P. “Estimation of Hydraulic Conductivity from Effective Porosity Data for
Some Indiana Soils,” Soil Science Society of America Journal, 55: 1801-1803, 1991.
Gerke, H.H., and M.Th. van Genuchten, “A Dual-Porosity Model for Simulating the
Preferential Movement of Water and Solutes in Structured Porous Media,” Water
Resources Research, 29: 305-319, 1993.
Good, P. I., 1999. Resampling methods: a practical guide to data analysis. Birkhäuser,
Boston.
Goovaerts, P. and C.N., Chiang, “Temporal Persistence of Spatial Patterns for
Mineralizable Nitrogen and Selected Soil Properties,” Soil Science Society of
America Journal, 57, 372-381, 1993.
Govindaraju, R.S. (ed.) “ Stochastic methods in subsurface contaminant
Grayson, R. B., and A.W. Western, “Towards Areal Estimation of Soil Water Content from
Point Measurements: Time and Space Persistence of Mean Response,” Journal of
Hydrology, 207: 68-82, 1998.
Griffioen, J.W., Barry, D.A., and J.Y. Parlange, “Interpretation of Two-Region Model
Parameters,”Water Resources Reserach, 34, 373-384, 1998.
Guadagnini, A., and Neuman, S.P., “Nonlocal and localized analyses of conditional mean
steady state flow in bounded randomly nonuniform domains 1. Theory and
computational approach,”Water Resour. Res. 35:2999–3018, 1999.
Guber, A. K., Rawls, W. J, Shein, E. V., and Y. A. Pachepsky, “Effect of Soil Aggregate
Size Distribution on Water Retention,” Soil Science, 168: 223 – 233, 2003.
Guber, A. K., Shein, E.V., Van Itsyuan and A.B.Umarova, “Experimental Information for
Mathematical Models of Water Transport in Soil: Assessment of Model Adequacy
and Reliability of Prediction,” Eurasian Soil Science, 31(9): 1020-1029, 1998.
Gupta, S. C., and W. E. Larson, “Estimating Soil Water Retention Characteristics from
Particle-Size Distribution, Organic Matter Percent, and Bulk Density,” Water
Resources Research, 15(6): 1633-1635, 1979.
Guswa, A. J. and D. L. Freuberg, “On Using the Equivalent Conductivity to Characterize
Solute Spreading in Environments with Low-Permeability Lenses,” Water
Resources Research, 38(8): 1132 (DOI 10.1029/2001WR000528), 2002.
Guswa, A. J., Celia, M .A., and I. Ridriguez-Iturbe, “Models of Soil Moisture In
Ecohydrology: A Comparative Study,” Water Resources Research, 38(9): 1166
(DOI 10.1029/2001WR000826), 2002.
Haise, H.R., Haas, H.J., and L.R. Jensen, “Soil Moisture of Some Great Plains Soils. II.
Field Capacity as Related To 1/3-Atmosphere Percentage and “Minimum Point” as
Related to 15- and 26-Atmosphere Percentages,” Soil Science Society of America
Proceedings, 19: 20-25, 1955.
124
Hall, D. G. M., Reeve, M. J., Thomasson, A. I., and V. F. Wright, “Water Retention,
Porosity And Density of Field Soils,” Soil Surv. Tech. Monogr. 9, 74 pp., 1977.
Hamill, T. M., Hansen, J. A., Mullen, S. L., Snyder, C. Workshop on Ensemble Forecasting
in the Short to Medium Range. Meeting Summary. Bulletin of the American
Meteorological Society, 86, (in press), 2005.
Hassan, A. and K. Hamed, “Prediction of Plume Migration in Heterogeneous Media Using
Artificial Neural Networks,” Water Resources Research, 37(3): 605-624
Hecht-Nielsen, R. “Neurocomputing,” Addison-Wesley Publishing Company, Reading,
MA, pp. 433 pp., 1990.
Heimovaara T.J. and W. Bouten, “A Computer Controlled 36-Channel Time Domain
Reflectometry System for Monitoring Soil Water Contents,” Water Resources
Research, 26: 2311-2316, 1990.
Hill, C. M. “Methods and Guidelines for Effective Model Calibration,” U. S. Geological
Survey Water Resources Investigations Report 98-4005, Denver, CO, 90 pp., 1998.
Hill, M. C., Middlemis, H., Hulme, P., Poeter, E., Riegger, J., Neuman, S. P., WIlliams, H.,
and M. Anderson, “Brief Overview of Selected Groundwater Modelling
Guidelines,” pp. 105-120, in: K. Kovar, J. Bruthans and Z. Hrkal (eds.), FiniteElement Models, MODFLOW, and More, Solving groundwater problems,
Proceedings, Carlsbad, Czech Republic, September 13-16, 2004.
Hillestad, R. J. and M. L. Juncosa, “Cutting Some Trees to Save the Forest: On
Aggregation and Disaggregation in Combat Models”, MR-189, RAND, Santa
Monica, Calif., 1993.
Hillestad, R., Owen, J. G. and Blumenthal D. Experiments in Variable Resolution
Modeling, N-3631-DARPA, RAND, Santa Monica, Calif., 1992.
Holder, M., Brown, K. W., Thomas, J. C., Zabcik, D. and H. E. Murray, “Capillary-Wick
Unsaturated Zone Soil Pore Water Samplers,” Soil Science Society of America
Journal, 55: 1195-1202, 1991.
Hubbard, S., J. Chen, J. Peterson, K. Williams, D. Watson, P. Jardine, and M. Fienen.
“Characterization & Monitoring at FRC Area 3 using Geophysical Data Conditioned
to Hydrogeological Data” at public.ornl.gov/nabirfrc/update_aug03.pdf , 2003
Hunt, R. and C. Zheng, “Debating Complexity in Modeling,” EOS, Transactions, American
Geophysical Union, 80(3): 29, January 19, 1999
Hupet, F. and M. Vanclooster, “Intraseasonal Dynamics of Soil Moisture Variability
Within a Small Agricultural Maize Cropped Field,” Journal of Hydrology, 261, 86–
101, 2002.
Innis, G., and E. Rextad, “Simulation model simplification techniques,” Simulation, 41: 715, 1983.
Iwasaki, Y., “Two Model Abstraction Techniques Based on Temporal Grain Size:
Aggregation and Mixed Models,” Knowledge Systems Laboratory Report, Stanford
University, 12 pp., 1990.
Iwasaki, Y., and H. Simon, “Causality and Model Abstraction,” Artificial Intelligence, 67:
143–194, 1992
Jacques D. “Analysis of Water Flow and Solute Transport at The Field Scale,” PhD, no
454, Faculteit Land bouwkundige en Toegepaste Biologische Wetenschappen, K.U.
Leuven, Belgium, 255 pp., 2000.
125
Jacques, D., Simunek, J., Timmerman, A., and J. Feyen, “Calibration of Richards and
Convection–Dispersion Equations To Field-Scale Water Flow And Solute Transport
Under Rainfall Conditions,” Journal of Hydrology, 259: 15-31, 2002.
Jacques, D., Timmerman, A. and J. Feyen, “Experimental Study of Water Flow and Solute
Transport in a Macroporous Soil Under Natural Boundary Conditions: BekkevoortSite Description, Experimental Design and Calibration Procedures,” Internal Report,
no. 53, Institute for Land and Water Management, K.U.Leuven, Belgium, 99 p.,
1999.
Jakeman, A.J. and G.M. Hornberger, “How Much Complexity is Warranted in a RainfallRunoff Model?” Water Resources Research, 29(8): 2637-2649, 1993.
Jarvis, N.J. “Modeling the Impact of Preferential Flow on Nonpoint Source Pollution,” pp.
195-221 in: H.M. Selim and L. Ma (eds.), Physical Nonequilibrium in Soils:
Modeling and Applications, Ann Arbor Press, Chelsea, MI. 1999.
Jothityangkoon, C., Sivapalan, M., and D.L. Farmer, “Process Controls of Water Balance
Variability in a Large Semi-Arid Catchment: Downward Approach To Hydrological
Modeling,” Journal of Hydrology 254: 174-198, 2001.
Jury, W.A. “Transfer Function Model of Field-Scale Solute Transport Under Transient
Water Flow,” Soil Science Society of America Journal, 54: 327-332, 1990.
Kachanoski, R.G. and E. De Jong, “Scale Dependence and the Temporal Persistence of
Spatial Patterns Of Soil Water Storage,” Water Resources Research, 24: 85-91,
1988.
Karpouzos, D. K., Delay, F., Katsifarakis, K. L., and G. de Marsily, “A Multipopulation
Genetic Algorithm to Solve the Inverse Problem in Hydrogeology. Water Resource
Research, 37: 2291-2302: 2001.
Kekkonen, T. S., and A. J. Jakeman, “A Comparison of Metric and Conceptual Approaches
in Rainfall Runoff Modeling and Its Implications,” Water Resources Research, 34:
2345-2352, 2001.
Kelson, V.A., Hunt, R.J., and H.M. Haitjema, “Improving a Regional Model Using
Reduced Complexity and Parameter Estimation,” Ground Water, 40(2): 132-143,
2002.
Kern, J. S. “Evaluation of Soil Water Retention Models Based on Basic Soil Physical
Properties,” Soil Science Society of America Journal, 59: 1134-1141, 1995.
Khaleel., R., Yeh, T. C. and Z. Lu, “Upscaled Flow and Transport Properties for
Heterogeneous Unsaturated Media,” Water Resources Research, 38(5): 1053 (DOI
10.1029/2000WR000072), 2002.
Knutson, J. H., Lee, S. B., Zhang, W. Q. and J. S Selker, “Fiberglass Wick Preparation for
Use on Passive Capillary Wick Soil Pore-Water Samplers,” Soil Science Society of
America Journal, 57: 1474-1476, 1993.
Koch, S., and H. Flühler. "Solute transport in aggregated porous media: Comparing model
independent and dependent parameter estimation", Water, Air, & Soil Pollution, 68:
275-289, 1993.
Kosugi, K.,”Three-Parameter Lognormal Distribution Model for Soil Water Retention,”
Water Resource Research, 30: 891-901, 1994.
Kron T. D. and D. Rosberg. “An Artificial Neural Network Based Groundwater Flow and
Transport Simulator,” http://www.isva.dtu.dk/staff/tdk/nn.2.pdf ,1998.
126
Kumar, P., “Layer Averaged Richard's Equation with Lateral Flow”, Advances in Water
Resources 27: 521 –531, 2004
Leenhardt, D., Voltz, M. and S. Rambal, “A Survey of Several Agroclimatic Soil Water
Balance Models with Reference to Their Spatial Application,” Eur. J. Agron., 4(1):
1-14, 1995.
Leij, F., W. J. Alves, M. Th. van Genuchten, and J. R. Williams, “The UNSODA
Unsaturated Soil Hydraulic Database. User's Manual Version 1.0” EPA/600/R96/095, National Risk Management Laboratory, Office of Research and
Development, Cincinnati, OH, 103 pp., 1996.
Li, S. G., McLaughlin D., and H. S. Liao, H. S., “A computationally practical method for
stochastic groundwater modelIng”, Adv. Water Resour. 26(11), 1137-1148, 2003.
Loague, K., “Using Soil Texture to Estimate Saturated Hydraulic Conductivity and the
Impact on Raifall-Runoff Simulations,” Water Resources Bulletin, 28: 687-693,
1992.
Mallants, D., Mohanty, B. Jacques, D. and J. Feyen, “Spatial Variability of Hydraulic
Properties in a Multi-Layered Soil,” Soil Science, 161:167-181, 1996.
Mallat, S. G. “A Wavelet Tour of Signal Processing,” Academic Press, San Diego, 673 pp.,
1998.
Marion, J. M., Or, D. Rolston, D. E., Kavvas, M. I and J. W. Biggar, “Evaluation of
Methods for Determining Soil-Water Retentivity and Unsaturated Hydraulic
Conductivity,” Soil Science, 158: 1-13, 1994.
Martinez-Fernandez, J. and A. Ceballos, “Temporal Stability of Soil Moisture in a LargeField Experiment in Spain,” Soil Science Society of America Journal, 67: 1647 –
1656, 2003.
Masket, S, Dibike, Y.B, Jonoski, A and D.P. Solomatine, “Groundwater Model
Approximation With ANN for Selecting Optimum Pumping Strategy for Plume
Removal,” pp. 67-80 in O. Schleider and A. Zijderveld, (eds.), AI Methods in Civil
Engineering Applications2nd Joint Workshop on AI Methods in Civil Engineering
Applications, Cottobus, 2000.
Mayr, T. and N. J. Jarvis, “Pedotransfer Functions to Estimate Soil Water Retention
Parameters for a Modified Brooks-Corey Type Model,” Geoderma, 91: 1-9. 1999.
McLaughlin D. “Recent Developments in Hydrologic Data Assimilation” Reviews of
Geophysics. Supplement: 977-984, US National report to International Union of
Geodesy and Geophysics, Paper 95RG00740, American Geophysical Union.
Washington, D. C., 1995.
Meisel, W. S. and D. C. Collins, “ Repromodeling: An approach to efficient model
utilization and interpretation’” IEEE Transactions on Systems, Man, and
Cybernetics SMC-3: 349-358, 1973.
Messing, I., “Estimation of Saturated Hydraulic Conductivity in Clay Soils from Moisture
Retention Data,” Soil Science Society of America Journal, 53: 665-668, 1989.
Meyer, P. D., Rockhold, M. L., Gee, G. W., and T. J. Nicholson, “Uncertainty Analyses of
Infiltration and Subsurface Flow and Transport for SDMP Sites,” NUREG/CR-6565,
PNNL-11705. U. S. Nuclear Regulatory Commission. Washington, DC 20555-0001,
1997.
Meyer, P.D., Taira, R. Y., and T. J. Nicholson, Hydrologic Uncertainty Assessment for
Decommissioning Sites: Hypothetical Test Case Applications. Pacific Northwest
127
National Laboratory, Richland, WA 99352. NUREG/CR-6695. Prepared for
Division of Risk Analysis and Applications Office of Nuclear Regulatory Research
U. S. Nuclear Regulatory Commission Washington, DC 20555-0001 NRC JCN
W6933, 34 pp., 2001.
Milly, P.C.D., “Climate, Soil Water Storage, And The Annual Average Water Balance,”
Water Resources Research, 30: 2143-2156, 1994.
Minasny, B. and Perfect, E., “Solute adsorption and transport parameters”, Pp. 195-224, In:
In: Pachepsky, Y. A., and Rawls, W J. (ed.). Development of pedotransfer functions
in soil hydrology. Elsevier, 512 pp., 2004.
Minasny, B., and. McBratney, A.B, “The Neuro-M Method for Fitting Neural Network
Parametric Pedotransfer Functions,” Soil Science Society of America Journal, 66: 352362, 2002.
Mitchell, M., “An Introduction to Genetic Algorithms,” MIT Press, Cambridge, MA, 205
p., 1996.
Mohanty, B.P., Bowman R.S., Hendrickx J.M.H., and M.Th.van Genuchten, “New
Piecewise-Continuous Hydraulic Functions for Modeling Preferential Flow in an
Intermittent Flood-Irrigated Field,” Water Resources Research, 33: 2049-2063,
1997.
Molteni, F., Buizza, R., Palmer, T.N., and Petroliagis, T. 1996: The ECMWF ensemble
prediction system: methodology and validation. Q. J. R. Met. Soc., 122, 73-119.
Nayak, P. “Causal Approximations,” Artificial Intelligence, 70: 277-334, 1994.
Neuman, S. and V. Di Federico, “Multifaceted Nature of Hydrogeologic Scaling and its
Interpretation,” Reviews of Geophysics, 41(3): 1014 (10.1029/2003RG000130),
2003
Neuman, S. P. “Universal Scaling of Hydraulic Conductivities and Dispersivities in
Geologic Media,” Water Resources Research, 26(8): 1749-1758, 1990.
Neuman, S. P., Wierenga, P. J., and T. J. Nicholson, “A Comprehensive Strategy of
Hydrogeologic Modeling and Uncertainty Analysis for Nuclear Facilities and Sites,”
NUREG/CR 6805. U. S. Nuclear Regulatory Commission. Washington, D.C. 205550001, 2003.
Neuman, S.P. “On Advective Dispersion Fractal Velocity in Permeability Fields,” Water
Resources Research, 31(6): 1455-1460. 1995.
Neupauer, R.M.and Powell, K.L., “A fully-anisotropic Morlet wavelet to identify dominant
orientations in a porous medium,” Computers & Geosciences (in press), 2005.
Nielsen, D.R., M. Th. Van Genuchten, and J. W. Biggar. “Water Flow and Solute Transport
Processes in the Unsaturated Zone,” Water Resources Research, 22 (9,suppl.): 89S108S, 1986.
Nikolaeva, S. A., Pachepsky, Ya. Shcherbakov, R. A. and A. I. Shcheglov. “Modeling
Moisture Regime of Ordinary Chernozems,” Pochvovedenie, 6: 52-59, 1986 (in
Russian).
Nimmo, J.R. “What Measurable Properties Can Predict Preferential Flow?” Paper No. 324.
Transactions of the 17th World Congress of Soil Science, August 14-21, Bangkok,
Thailand, 2002.
Oliver, Y.M., and K.R.J. Smettem, “Parameterization of Physically Based Solute Transport
Models In Sandy Soils,” Aust. J. Soil Res., 41: 771-788, 2003.
128
Oosterveld, M., and C. Chang, “Empirical Relations Between Laboratory Determinations
of Soil Texture and Moisture Characteristic,” Can. Agric. Eng. 22:149-151, 1980.
Pachepsky, Y. A., Guber, A. K., and D. Jacques, “Temporal Persistence in Vertical
Distributions of Soil Moisture Contents,” Soil Science Society of America Journal,
69(2) (in press), 2005.
Pachepsky, Y. A., Timlin, D. J., and W. J. Rawls, Generalized Richards’ Equation to
Simulate Water Transport in Unsaturated Soils,” Journal of Hydrology, 272: 3–13,
2003.
Pachepsky, Y., Benson, D. and W.J. Rawls, “Simulating Scale-Dependent Solute Transport
in Soils With the Fractional Advective-Dispersive Equation,” Soil Science Society of
America Journal, 64:1234-1243, 2000.
Pachepsky, Y.A., and Rawls, W. J. (eds.), “Development of Pedotransfer Functions in Soil
Hydrology,” Developments in Soil Science, 30, Amsterdam, Elsevier, 2004.
Pachepsky, Y.A., and W.J. Rawls, “Soil Structure and Pedotransfer Functions,” European
Journal of Soil Science, 54: 443-452, 2003.
Pachepsky, Y.A., Mironenko, E.V., and R.A. Scherbakov, “Prediction and Use of Soil
Hydraulic Properties,” pp. 203-212 in: M.Th. van Genuchten, F.J. Leij, and L.J.
Lund (eds.), Indirect Methods for Estimating the Hydraulic Properties of
Unsaturated Soils, Riverside: University of California, 1992.
Pachepsky, Y.A., Timlin, D.J., and W.J. Rawls, “Soil Water Retention as Related to
Topographic Variables,” Soil Science Society of America Journal, 65: 1787-1795,
2001.
Pachepsky, Ya. , Benson D. and W. Rawls, “Simulating Scale-Dependent Solute Transport
in Soils With the Fractional Advective-Dispersive Equation,” Soil Science Society of
America Journal, 64: 1234-1243, 2000.
Pachepsky, Ya. A., and W. J. Rawls, “Accuracy And Reliability of Pedotransfer Functions
as Affected by Grouping Soils,” Soil Science Society of America Journal, 63: 17481757, 1999.
Pachepsky, Ya. A., Shcherbakov, R. A., Varallyay, Gy. and K.Rajkai, “Statistical Analysis
of the Correlation Between the Water Retention and other Physical Soil Properties,”
Pochvovedenie 2: 56-66 (in Russian), 1982.
Pachepsky, Ya. A., W. J. Rawls, and D. J. Timlin, “The Current Status of Pedotransfer
Functions: Their Accuracy, Reliability, and Utility In Field- and Regional-Scale
Modeling,” pp. 223-234 in D. L.Corwin, , K. Loague, and T. R. Ellsworth (Eds.),
Assessment of Non-point Source Pollution in the Vadose Zone, Geophysical
monograph 108. American Geophysical Union, Washington, D. C., 1999.
Pachepsky, Ya., D. Timlin, and L. Ahuja. “Estimating Saturated Soil Hydraulic
Conductivity Using Water Retention Data and Neural Networks,” Soil Science,
164:552-560, 1999.
Palmer, R.N. and J.L. Cohan, “Complexity in Columbia River Systems Modeling.” Journal
of Water Resources Planning and Management-ASCE, 112(4): 453-468, 1986.
Palmer, T.N., A. Alessandri, U. Andersen, P. Cantelaube, M. Davey, P. Délécluse, M.
Déqué, E. Díez, F.J. Doblas-Reyes, H. Feddersen, R. Graham, S. Gualdi, J.-F.
Guérémy, R. Hagedorn, M. Hoshen, N. Keenlyside, M. Latif, A. Lazar, E.
Maisonnave, V. Marletto, A. P. Morse, B. Orfila, P. Rogel, J.-M. Terres, M. C.
Thomson, Development of a European multi-model ensemble system for seasonal to
129
inter-annual prediction (DEMETER). Bulletin of the American Meteorological
Society, 85, 853-872, 2004.
Pan, L. and P.J. Wierenga, “A Transformed Pressure Head Based Approach to Solve
Richards' Equation for Variably Saturated Soils,” Water Resources Research, 31:
925-931, 1995.
Perfect, E., McLaughlin, N. B., Kay, B. D. and G. C. Topp, “An Improved Fractal Equation
for the Soil Water Retention Curve,” Water Resources Research, 32: 281-287, 1996.
Petersen, G. W., Cunningham, R. L., and R. P. Matelski, “Moisture Characteristics of
Pennsylvania Soils: I. Moisture Retention as Related To Texture,” Soil Science
Society of America Proceedings, 32: 271-275, 1968.
Pohlman, K. F., Hassan, A. E., Chapman, J. B. “Modeling Density-Driven Flows and
Radiounuclide Transport at an Underground Nuclear Test: Uncertainty Analysis and
Effect of Parameter Correlation,” Water Resource Research, 385): 1059 (DOI
10.29/2001WR001047), 2002.
Pruess, K. “A General-Purpose Numerical Simulator for Multiphase Fluid and Heat Flow,”
Report LBL-29400, Lawrence Berkeley Laboratory, Berkeley, CA, 103 p., 1991.
Rajkai, K. and Gy. Várallyay, “Estimating Soil Water Retention from Simpler Properties by
Regression Techniques,” pp.417-426 in M. Th. van Genuchten, F. J. Leij, and L. 3.
Lund (eds.), Proc. Int. Workshop on Indirect Methods for Estimating the Hydraulic
Properties of Unsaturated Soils, University of California, Riverside, CA, 1992.
Rawls, W. J., and D. L. Brakensiek, “Prediction of Soil Water Properties for Hydrologic
Modeling,” pp.293-299 in E. B. Jones and T. J. Ward (eds.), Proc. Symp. Watershed
Management in the Eighties, April 30-May 1, 1985, Denver, CO. Am. Soc. Civil
Eng., New York, NY. 1985.
Rawls, W. J., and D. L. Brakensiek. 1989. Estimation of Soil Water Retention and
Hydraulic P roperties. p.275-300. In Morel-Seytoux (ed.) Unsaturated Flow in
Hydrologic Modeling; Theory and Practice. Proc. NATO Adv. Res. Workshop
Hydrology,” ASCE Press, New York, 2002.
ASI Series, Ser. C, Vol.275. KIuwer Academic Press, Dordrecht, The Netherlands.
Rawls, W. J., and Ya. A. Pachepsky, “Using Field Topographic Descriptors to Estimate
Soil Water Retention,” Soil Science, 167: 423-235, 2002
Rawls, W. J., Brakensiek, D. L., and B. Soni, “Agricultural Management Effects on Soil
Water Processes, Part I. Soil Water Retention and Green-Ampt Parameters,” Trans.
ASAE 26: 1747-1752, 1983.
Rawls, W. J., Brakensiek, D. L., and K. E. Saxton, “Estimation of Soil Water Properties,”
Trans. ASAE 25: 1316-1320, 1982.
Rawls, W. J., L R. Ahuja, and D. L Brakensiek. “Estimating Soil Hydraulic Properties
from Soils Data,” p.329-340 in M. Th. van Genuchten, F. J. Leij, and L. Lund (eds.)
Proc. Int. Workshop on Indirect Methods for Estimating the Hydraulic Properties of
Unsaturated Soils, University of California, Riverside, CA. 1992.
Rawls, W.J., Giménez, D., and Grossman, R., “Use of soil texture, bulk density and slope
of the water retention curve to predict saturated hydraulic conductivity. Trans.
ASAE 41(4): 983-988, 1998.
130
Reichardt, K., Bacchi, O.O.S., Villagra, M.M., Turatti, A.L. and Z.O., Pedrosa, “Hydraulic
Variability in Space and Time in a Dark Red Latosol of The Tropics,” Geoderma,
60: 159-168, 1993.
Reilly, T. E. and A. W. Harbaugh, “Guidelines for Evaluating Ground-Water Flow
Models,” U. S. Geological Survey Scientific Investigation Report 2004-5038, 30 p.,
2004.
Reynolds, S.G., “The Gravimetric Method of Soil Moisture Determination, Part III: An
Examination of Factors Influencing Soil Moisture Variability,” Journal of
Hydrology, 11: 288-300, 1970.
Rickel, J., and P. Porter, ``Automated Modeling of Complex Systems to Answer Prediction
Questions'', Artificial Intelligence Journal, 93(1-2), pp. 201--260, 1997.
Rivers, E.D. and R.F. Shipp, “Available Water Capacity of Sandy and Gravely North
Dakota Soils,” Soil Science, 113: 74-80,1972.
Romano, N., and M. Palladino, “Prediction of Soil Water Retention Using Soil Physical
Data and Terrain Attributes,” Journal of Hydrology., 265: 56-75, 2002.
Rubin, Y., “Applied stochastic hydrogeology,”Oxford University Press, Oxford, 2003
Saxton, K. E., W. J. Rawls, J. S. Romberger, and R. I. Papendick, “Estimating Generalized
Soil-Water Characteristics from Texture,” Soil Science Society of America Journal,
50:1031-1036, 1986.
Scanlon, B. R., Tyler, S. W., P. J. Wierenga, “Hydrologic Issues in Arid, Unsaturated
Systems And Implications for Contaminant Transport,” Rev. Geophys., 33: 461-490,
1997.
Scanlon, B. Software Review, 2004.
http://typhoon.mines.edu/software/igwmcsoft/hydrus1d _review.htm
Schaap, M.G. “Accuracy and uncertainty of PTF predictions”. Pp. 33-42. In: Pachepsky, Y.
A., and Rawls, W J. (ed.). Development of pedotransfer functions in soil hydrology.
Elsevier, 512 pp., 2004.
Schaap, M.G. and F. J. Leij. “Database-related Accuracy and Uncertainty of Pedotransfer
Functions,” Soil Science, 163: 765-779, 1998.
Schaap, M.G. and F.J. Leij, “Improved Prediction of Unsaturated Hydraulic Conductivity
with the Mualem-Van Genuchten Model,” Soil Science Society of America Journal,
64:843-851, 2000.
Schaap, M.G. and F.J. Leij, “Improved Prediction of Unsaturated Hydraulic Conductivity
with the Mualem-Van Genuchten Model,” Soil Science Society of America Journal,
64:843-851, 2000.
Schaap, M.G., Leij, F.G., van Genuchten, M.T., “Neural network analysis for hierarchical
prediction of soil hydraulic properties”, Soil Sci. Soc. Am. J, 62:847-855, 1998.
Schaap, M.G., Leij, F.J., and M.Th. van Genuchten, “Rosetta: A Computer Program for
Estimating Soil Hydraulic Parameters with Hierarchical Pedotransfer Functions,”
Journal of Hydrology, 251:163-176, 2001.
Schmied, B., Abbaspour, K., and R. Schulin, “Inverse Estimation of Parameters in a
Nitrogen Model Using Field Data,” Soil Science Society of America Journal, 64:
533–542, 2000.
Simon, H. and A. Ando, “Aggregation of Variables in Dynamic Systems,” Econometrica,
29: 111–138, 1996.
131
Simunek, J., M. Sejna, and M. Th. van Genuchten, “The HYDRUS-1D software package
for simulating the one-dimensional movement of water, heat, and multiple solutes in
variably-saturated media. Version 2.0”, IGWMC - TPS - 70, International Ground
Water Modeling Center, Colorado School of Mines, Golden, Colorado, 186pp.,
1998
Sisty, A.F. and S.D. Farr, “Model Abstraction Techniques: An Intuitive Overview,” Air
Force Research Laboratory/IFSB, Rome, New York, 2004.
http://www.rl.af.mil/tech/papers/ ModSim/ModAb_Intuitive.html
Smettem, K.R.J., Oliver, Y.M., Pracilio, G., and R.J. Harper, “Data Availability and Scale
in Hydrologic Applications,” pp.253-272, in: Pachepsky, Y. and Rawls, W.J. (eds.),
Development of Pedotransfer Functions in Soil Hydrology. Amsterdam, Elsevier,
2004.
Smith, S.K., “Further Thoughts on Simplicity and Complexity in Population Projection
Models,” International Journal of Forecasting, 13(4): 557 – 565, 1997.
Soil Survey Staff, 1951. Soil Survey Manual. USDA Handbook No. 18. US Government
Printing Office. Washington, D.C.
Statsoft, “STATISTICA for Windows. Computer Program Manual”, StatSoft, Inc., Tulsa,
CD-ROM, 1999.
Stephens, P. A., Frey-Roos, F. Arnold, W., and W. J. Sutherland, “Model Complexity and
Population Predictions. The Alpine Marmot As A Case Study,” Journal of Animal
Ecology, 71: 343–361, 2002.
Stillman R. A. , McGrorty, S., Goss-Custard, J. D. and A. D. West, “Predicting Mussel
Population Density and Age Structure: The Relationship Between Model
Complexity and Predictive Power,” Marine Ecology Progress Series 208: 131–145,
2000
Suleiman, A.A. and J. T. Ritchie, “Modeling Soil Water Redistribution Under Second Stage
Evaporation,” Soil Science Society of America Journal, 67(2): 377-386, 2003.
Tamboli P.M., Larson, W.E. and M. Amemya, “Influence of Aggregate Size on Moisture
Retention,” Iowa Acad. Sci. 71: 103-108, 1964.
Therrien, R., and E.A. Sudicky, “Three-Dimensional Analysis of Variably-Saturated Flow
and Solute Transport in Discretely-Fractured Porous Media,” Journal of
Contaminant Hydrology, 23: 1-44, 1996.
Tietje, O., and M. Tapkenhinrichs. “Evaluation of Pedo-transfer Functions,” Soil Science
Society of America Journal, 57: 1088-1095, 1993.
Timlin, D. J., L. R. Ahuja, and R. D. Williams, “Soil Hydraulic Parameters for RegionalScale Applications of Mechanistic Models,” pp. 185- 203 in D. L. Corwin and K.
Loague (eds.), Application of GIS to the modeling of non-point source pollutants in
the vadose zone. SSSA Special publication 48, American Society of Agronomy,
Madison, WI, 1996b.
Timlin, D. J., Y. A. Pachepsky, and C. L. Walthall. “A Mix of Scales: Topography, Point
scales, and Yield Maps,” pp. 227-242, in Y. A. Pachepsky, D. E. Radcliffe, and H.
M. Selim (eds.), Scaling Methods in Soil Physics, CRC Press, Boca Raton, FL,
2003.
Timlin, D. J., Ya. Pachepsky, B. Acock, and F. Whisler. “Indirect Estimation of Soil
Hydraulic Properties to Predict Soybean Yield Using GLYCIM,” Agricultural
Systems, 52: 331-353, 1996a.
132
Tomasella, J. and M.G. Hodnett, “Estimating Soil Water Retention Characteristics from
Limited Data in Brazilian Amazonia,” Soil Science, 163: 190-202, 1998.
Tsai, F. T-C., N-Z. Sun, and W. W-G. Yeh, “Global-Local Optimization Methods for the
Identification of Three-Dimensional Parameter Structure in Groundwater
Modeling,” Water Resources Research, 39, 1043, doi:10.1029/2001WR001135,
2003
U.S. NRC, “Branch Technical Position on Site Characterization for Decommissioning,” U.
S. Nuclear Regulatory Commission, Washington, DC, 1994.
U.S. NRC, “Site Decommissioning Management Plan,” NUREG-1444, U. S. Nuclear
Regulatory Commission, Washington, DC, 1993.
Unsaturated Flow in Hydrologic Modelling. ArIes, France, June 13-17 1988. NATO
Vachaud, G., Passerat de Silans, A., Balabanis, P., and M. Vauclin, “Temporal Persistence
of Spatially Measured Soil Water Probability Density Function,” Soil Science
Society of America Journal, 49: 822- 828, 1985.
Valocchi, A.J. “Validity of The Local Equilibrium Assumption For Modeling Sorbing
Solute Transport Through Homogeneous Soils,” Water Resources Research, 21:
808-820, 1985.
Van Dam, J.C. and R.A. Feddes, “Numerical Simulation of Infiltration, Evaporation and
Shallow Groundwater Levels with the Richards Equation,” Journal of Hydrology,
233: 72-85, 2000.
Van Genuchten, M. T., Leij, F. J., and Lund, L. J. (eds.), “Proceedings of the International
Workshop on Indirect methods for Estimating the Hydraulic Properties of
Unsaturated Soils”, University of California, Riverside, CA, 1992.
Van Genuchten, M. Th, F. J. Leij, and L. Wu ( Eds.). “Characterization and Measurement
of the Hydraulic Properties of Unsaturated Porous Media,” Proceedings of the
International Workshop, Riverside, California, October, 22-24, 1997, University of
California, Riverside, 1997.
Van Genuchten, M. Th. “A Closed Form Equation for Predicting the Hydraulic
Conductivity Of Unsaturated Soils,” Soil Science Society of America Journal, 44:
892-898, 1980.
Van Genuchten, M.Th., “Non-equilibrium transport parameters from miscible displacement
experiments. Research Rep. No. 119,” U.S. Salinity Laboratory, USDA-SEA-ARS,
Riverside, CA, 1981.
Van Ness, E. H., Scheffer, M, “A strategy to improve the contribution of complex
simulation models to ecological theory”, Ecological Modelling, 185: 153-164, 2005.
Ventrella, D., Mohanty, B. P., Simunek, J. Losavio, N., and M. Th. van Genuchten, “Water
and Chloride Transport in a Fine-Textured Soil: Field Experiments and Modeling,”
Soil Science, 165(8): 624-631, 2000.
Vereckeen, H., Maes, J., Feyen, J., and P. Darius, “Estimating the Soil Moisture Retention
Characteristics from Texture, Bulk Density and Carbon Content,” Soil Science, 148:
389-403, 1989.
Wang, C. G., and D. G. Jamieson, “An Objective Approach to Regional Wastewater
Treatment Planning,” Water Resource Research, 38(3): 1022, (DOI
10.1029/2000WR000062), 2002.
133
Wang, W, Neuman, S. P., Yao, T., and P. J. Wierenga, “Simulation of Large-Scale Field
Infiltration Experiments Using a Hierarchy of Models Based on Public, Generic, and
Site Data,” Vadose Zone Journal, 2: 297–312, 2003.
Weiner, E. S. C. and Simpson, J. A. (Eds.). “The Compact Oxford English Dictionary,”
Oxford University Press, 1991.
Weld, D. “The Use of Aggregation in Causal Simulation,” Artificial Intelligence, 30: 47-55,
1986.
Whelan, G., J. P. McDonald, and C. Sato, “Multimedia Environmental Pollutant
Assessment System (MEPAS): Groundwater Pathway Formulations,” PNNL-10907,
Pacific Northwest National Laboratory, Richland, WA, 1996.
Whitmore, A.P., “A method for assessing the goodness of computer simulation of soil
processes,” J. Soil Sci., 42:289-299, 1991.
Williams, J., Ross, P., and K. Bristow, “Prediction of the Campbell Water Retention
Function From Texture, Structure, And Organic Matter,” pp. 427-442 in M. Th. van
Genuchten, F. J. Leij, and L J. Lund (eds.), Proceedings of the International
Workshop on Indirect methods for Estimating the Hydraulic Properties of
Unsaturated Soils, University of California, Riverside, CA, 1992.
Wilson, G.V., Gwo, J.P., Jardine, P.M., and R.J. Luxmoore, “Hydraulic and
Nonequilibrium Effects on Multiregion Flow,” pp. 37-61 in: H.M. Selim and L. Ma
(eds.), Physical Nonequilibrium in Soils; Modeling and Application, Ann Arbor
Press, Chelsea, MI, 1998.
Wittmuss, H.D., and A.P. Mazurak, “Physical and Chemical Properties of Aggregates in a
Brunizem Soil,” Soil Science Society of America Proceedings, 22: 1-5, 1958.
Wösten J.H.M., A. Lilly, A. Nemes, and C. Le Bas, “Development and use of a database
of hydraulic properties of European soils,” Geoderma, 90, 169-185, 1999.
Wösten, J.H.M., Y. A. Pachepsky, and W. J. Rawls, “Pedotransfer Functions: Bridging the
Gap Between Available Basic Soil Data and Missing Soil Hydraulic
Characteristics,” Journal of Hydrology, 251: 123-150, 2001.
Wu, Y.-S., Zhang, K., and K. Pruess, “Massively Parallel Simulation of Flow and Transport
in Variably Saturated Porous and Fractured Media,” Lawrence Berkeley National
Laboratory. Paper LBNL-49407, 8 pp., 2002.
Ye, M., Neuman, S. P., and Meyer, P. D., ”Maximum likelihood Bayesian averaging of
spatial variability models in unsaturated fractured tuff,”. Water Resour. Res., 40,
W05113, doi:10.1029/2003WR002557, 2004
Yi-Bing Lin and P. A. Fishwick, “Asynchronous Parallel Discrete Event Simulation,” IEEE
Transactions on Systems, Man and Cybernetics, 26: 397-412, 1996.
Yilmaz, L. and T.I. Ören, Dynamic “Model Updating in Simulation with Multimodels: A
Taxonomy and a Generic Agent-Based Architecture,” Proceedings of SCSC 2004 Summer Computer Simulation Conference, July 25-29, 2004, San Jose, CA., pp. 3-8.
2004.
Zeigler, B. “Theory of Modeling and Simulation,” Academic Press, NY, 510 pp., 2000.
Zeigler, B. P., Kim, T. G., and Praehofer, H, “Theory of Modeling and Simulation. 2 ed.,
New York, NY, Academic Press, 1998.
Zeigler, B., “Theory of modeling and simulation,” New York, New York: Wiley and Sons,
1976.
134
Zhang, T., and R. Berndtsson, “Temporal Patterns and Spatial Scale of Soil Water
Variability In a Small Humid Catchment,” Journal of Hydrology, 104: 111-128,
1988.
Zhu, J. and B. P. Mohanty, “Spatial Averaging of Van Genuchten Hydraulic Parameters for
Steady State Flow in Heterogeneous Soils: A Numerical Study,” Vadose Zone
Journal, 1: 261-272, 2002.
Zurmühl, T. “Capability Of Convection-Dispersion Transport Models To Predict Transient
Water and Solute Movement in Undisturbed Soil Columns,” Journal of
Contaminant Hydrology, 30: 101-110, 1998.
Zurmühl, T., and W. Durner, “Modeling Transient Water and Solute Transport in a
Biporous Soil,” Water Resources Research, 32: 819-829, 1996.
135
Appendix A. Pedotransfer functions to estimate soil water retention
Most of equations presented below were developed from large (>200) samples databases.
Symbols w and θ denote volumetric (cm3cm-3) and gravimetric (cm3 g-1) water contents,
respectively, subscripts indicate the capillary pressure in cm, clay, and sand denote
percentages of textural fractions according the USDA textural classification, OM is the
organic matter content, %, OC is the organic carbon content, ρb is the bulk density g cm-3,
other symbols are defined as they appear.
A.1. Equations developed to estimate soil water content at fixed soil water potentials
Baumer (1992) used the US National Soil Survey database to relate gravimetric water
contents at capillary pressures of 15000 cm and 330 cm with the equations:
w15000=0.01*ρb*(0.71+0.45*OM+0.336*clay+0.117*clay*(CA3/2)+0.004*clay*SAR) (A1)
w330=0.01*ρb*(15.84+0.746*OM+2.2025*CA2(A2)
0.137*sand+0.743*w15000)
if clay > 10%
w330=0.01*ρb*(15.84+0.746*OM+0.02*CA2*clay2(A3)
0.137*sand+0.743*w15000) if clay ≤ 10%
where CA is the clay activity, SAR is the sodium adsorption ratio.
Bruand et al. (1994) estimated volumetric water contents at capillary pressures of 15000
cm and 330 cm as
θ15000 = (0.008+0.00367*clay)/(0.471+0.00411*clay)
(A4)
θ330 = (0.043+0.004*clay)/(0.471+0.00411*clay)
(A5)
Canarache (1993) applied regression analysis to the Romanian national database and
obtained predictive equations:
θ15000=0.01*ρb*(0.2805*clay+0.0009615*clay2)
(A6)
2
3
θ330=0.01*ρb*(2.65+1.105*clay-0.01896*clay +0.0001678*clay +15.12*ρb-6.745*ρb2(A7)
0.1975*clay*ρb)
Gupta and Larson (1979) used a subset of the US National Cooperative Survey database
to derive equations for estimating volumetric water contents at capillary pressures of
15000 cm and 330 cm as
θ330=0.003075*sand+0.005886*silt+0.008039*clay+0.002208*OM-0.1434*ρb
(A8)
θ15000=-0.000059*sand+0.001142*silt+0.005766*clay+0.00228*OM+0.02671*ρb (A9)
Hall et al. (1977) analyzed a subset of British Soil Survey data and arrived to the
equations
θ330=0.01*(20.81+0.45*clay+0.13*silt-5.95*ρb)
(A10)
θ15000=0.01*(1.48+0.84*clay-0.0055*clay2)
(A11)
Petersen et al. (1968) worked with the Pennsylvania soil database. Their equations are:
θ330=0.01*(11.83+0.96*clay-0.008*(clay)2)
(A12)
θ15000=0.01*(1.74+0.76*clay –0.005(clay)2)
(A13)
A-1
Rajkai and Varallyay (1992) analyzed the Hungarian national database:
θ330=0.01*(38.62-0.00479*sand-0.0019*(sand/silt)2)
θ15000=0.01*(1.39+0.36*clay +0.22*OM2)
We used clay content instead of unavailable fine fraction content in Eq. (A14).
(A14)
(A15)
Tomasella and Hodnett (1998) have been working with Brazilian soils and came up with
the equations:
θ330=0.01*(4.046+0.426*silt+0.404*clay)
(A16)
θ15000=0.01*(0.91+0.150*silt+0.396*clay)
(A17)
A.1.2 Equations developed to estimate Brooks-Corey parameters
Rawls and Brakensiek (1989) developed the following equations to estimate the BrooksCorey parameters in Eq. (1)
hb =exp[5.340+0.185*clay-2.484*φ -0.002*clay2-0.044*sandφ +0.001*sand2*φ 20.009*clay2*φ 2 -0.00001*sand2*clay+0.009*clay2*sand-0.0007*sand2*φ
0.000005*clay2*sand-0.500*φ 2*clay]
(A18)
λ=exp[-0.784+0.018*sand-1.062*φ -0.00005*sand2-0.003*clay2+1.111*φ 2-0.031*sand*φ
+0.0003*sand2*φ 2-0.006*clay2*φ 2-0.000002*sand2*clay+0.008*clay2*φ
-0.007*φ 2*clay]
(A19)
θr=-0.018+0.0009*sand+0.005*clay+0.029*φ -0.0002*clay2-0.001*sand*φ 0.0002*clay2*φ 2+0.0003*clay2*φ -0.002*φ 2*clay
(A20)
Campbell and Shiozava (1992) set the value of the residual water content in Eq. (1) equal
to zero and transformed this equation to
−b
h = he (θ / θ s )
(A21)
where he is “air entry potential”. The parameters in (1), estimated from two data sets for
British soils, were found to be
hes = −0.05d g−1 / 2
(A22)
b = −20hes + 0.2σ g
where the value of hes corresponds the air entry potential at a standard bulk density, ρb of
1.3 g cm-3. The proposed adjustment for bulk density is
0.67 b
he = hes (ρ b / 1.3)
(A23)
The geometric mean diameter ds and geometric standard deviation are given by
dg=exp(-0.025- 0.0363silt – 0.0688clay)
(A24)
2
1/2
σg=exp(0.133silt +0.477clay - ln dg)
Saxton et al. (1986) also set the value of the residual water content in Eq. (1) equal to
zero and transformed this equation to
h = Aθ B
(A25)
where
A=100*exp(-4.396-0.0715*clay-0.000488*sand2-0.00004285*sand2*clay)
(A26)
B=-3.140-0.00222*clay2-0.00003484*sand2*clay
(A27)
A-2
Oosterveldt & Chang (1980) transformed Eq. (1) with θr=0 to the form
θ = 0.01*ρb* (35.367+0.644*clay-0.251*sand-0.045*D) *h -0.190
where D is the mean depth of sample in centimeters.
(A28)
Williams et al. (1992) transformed Eq. (1) with θr=0 to the logarithmic form
lnθ =A+Blnh
(A29)
and applied it to the Australian database. Pedotransfer equations depended on the
availability of data on basic properties. Equations
A=2.57+0.238*ln(clay)-0.000192*sand2-0.0137*sand0.0926*ln(OM)+0.0412*OM
(A30)
B=-0.403+0.0871*ln(clay)-0.00077*sand
(A31)
were suggested for cases when the data on organic matter are available, and equations
A=1.839+0.257*ln(clay)+0.381*2-0.0001*sand2
(A32)
2
(A33)
B=-0.303+0.093*ln(ρb)+0.0565*ln(clay)-0.00003*sand
were developed or cases when no information about the organic matter content was
available.
Mayr and Jarvis (1999) set θr=0 and porosity φ equal to the saturated water θs in the
Brooks-Corey equation, and combined this equation in the dry range of soil water
retention curve
θ = θ s ( h / a ) −1 / b
θ < θi
(A34)
with a parabolic equation in the wet range:
θ s h 2 (1 − θ i / θ s )
(A35)
θ = θs − 2
θ ≥ θi
a (θ i / θ s ) − 2b
The water content θi and the equivalent pressure head hi at the matching point are given
by:
2bθ s
(A36)
θi =
1 + 2b
−b
⎛ 2b ⎞
hi = a⎜
(A37)
⎟
⎝ 1 + 2b ⎠
Pedotransfer functions developed from a Scandinavian dataset were:
log(a)=-4.9840297533+0.0509226283*sand+0.1575152771*silt+0.1240901644*ρb
-0.1640033143*OC-0.0021767278*silt2+0.0000143822*silt3+0.0008040715*clay2
+0.0044067117*OC2)
(A38)
log(1/b)=-0.8466880654-0.0046806123*sand+0.0092463819*silt-0.4542769707*ρb
-0.0497915563*OC+0.0003294687*sand2
(A39)
0.000001689056*sand3+0.0011225373*OC2
θs = 0.2345971971+0.0046614221*sand+0.0088163314*silt+0.0064338641*clay(A40)
0.3028160229*ρb +1.79762*102*sand2-3.134631*102*silt2
A.1.3 Equations developed to estimate Brooks-Corey parameters
A-3
Wösten et al. (1999) analyzed the all-Europe database and derived pedotransfer functions
to estimate van Genuchten parameters in the Eq. (2):
θs = 0.7919 + 0.001691*clay - 0.29619*ρb - 0.000001491*silt2 + 0.0000821*OM2 +
0.02427/clay + 0.01113/silt + 0.01472*ln(silt) - 0.0000733*OM*clay - 0.000619*ρb*clay (A41)
0.001183*ρb*OM - 0.0001664*topsoil*silt
α =exp[-14.96 + 0.03135*clay + 0.0351*silt + 0.646*OM +15.29*ρb - 0.192*topsoil 4.671*ρb2 - 0.000781*clay2 - 0.00687*OM2 + 0.0449/OM + 0.0663*ln(silt) +
0.1482*ln(OM) - 0.04546*ρb *silt - 0.4852*ρb*OM +
0.00673*topsoil*clay]
(A42)
2
n =1.+exp[-25.23 - 0.02195*clay + 0.0074*silt - 0.1940*OM + 45.5*ρb - 7.24*ρb +
0.0003658*clay2 + 0.002885*OM2 -12.81/*ρb - 0.1524/silt - 0.01958/OM - 0.2876*ln(silt)
- 0.0709*ln(OM) -44.6*ln(ρb) - 0.02264*ρb*clay + 0.0896*ρb*OM +
(A43)
0.00718*topsoil*clay
where topsoil is an ordinal variable having the value of 1 or of 0. Parameter m in Eq. (1)
was computed as 1-1/n.
These authors have also estimated average values of van Genuchten parameters by the
FAO textural classes. Those values are shown in Table A1 and the FAO textural classes
are delimited in Fig. A1.
Vereecken et al. (1989) used the Belgian dataset to develop the following pedotransfer
functions for van Genuchten equation with m=1:
θs = 0.81 – 0.283*ρb + 0.001*clay
(A44)
θr = 0.015 + 0.005*clay + 0.014*OC
(A45)
α= exp(-2.486 + 0.025*sand – 0.351*clay)
(A46)
(A47)
n=0.81 – 0.283*ρb + 0.001*clay
Pachepsky et al. (1982) applied van Genuchten equation with m=1 and θr = 0 to the
Hungarian national database and found regression equations:
θs = 0.01*(-56.4*ρb + 0.00205*clay2 + 123.79)
(A48)
(A49)
n = 0.336*ρb - 0.053
-0.04701*ρb*clay + 1.513*ρb – 0.417
(A50)
α= 10
A-4
Table A1. Tabulated parameters of van Genuchten equation by FAO textural classes
(Wösten et al., 1999)
n
FAO textural
θs
θr
α
class
Topsoils
Coarse
0.025 0.403 0.0383 1.3774
Medium
0.010 0.439 0.0314 1.1804
Medium fine 0.010 0.430 0.0083 1.2539
Fine
0.010 0.520 0.0367 1.1012
Very fine
0.010 0.614 0.0265 1.1033
Subsoils
Coarse
0.025 0.366 0.0430 1.5206
Medium
0.010 0.392 0.0249 1.1689
Medium fine 0.010 0.412 0.0082 1.2179
Fine
0.010 0.481 0.0198 1.0861
Very fine
0.010 0.538 0.0168 1.0730
A-5
Figure A1. Textural classification systems according to (a) USDA (Soil Survey Staff, 1951) and (b) as defined by the FAO Soil Map
of Europe (European Soil Bureau, 1998). USDA classes: S, sand, lS, loamy sand, sL, sandy loam, scL, sandy clay loam, sC, sandy
clay, L, loam, cL, clay loam, C, clay, siL, silt loam, sicL, silty clay loam, siC, silty clay, Si, silt; FAO classes, C, coarse, M, medium,
MF, medium fine, F, fine, VF, very fine.
A-6
A-1
Appendix B. Diurnal oscillations in water content and capillary
pressure data
Inspection of time series of water content and pressure head revealed oscillatory behavior
that apparently had diurnal periodicity. An example of such behavior is shown in Fig. B1
for the 10-day period between day 80 and day 90 from the beginning of the experiment at
location “10“. Both pressure head and water content (Fig. B1.a and Fig. B1.b) oscillated
synchronously at all depths. Temperature time series at different depths had sine wave
shape. The amplitude decreased and a phase had been shifted with depth (Fig. B1.c). Our
first hypothesis was that the oscillations of water contents and capillary pressures at each
depth could be related to temperature at this depth. To test this hypothesis, capillary
pressure and temperature time series were subjected to spectral Fourier analysis with the
STATISTICA software (StatSoft, 1999). Values of periods were equal to 24 hour for all
pressure head time series (Fig. B2.a), and for soil temperature time series at depth 15 cm
and 35 cm (Fig. B2.b). However, temperature time series at depths below 35 cm did not
have the 24-hour period. Therefore, local oscillations of temperature could not be useful
in explaining oscillations of capillary pressure and water contents at depths larger than 35
cm. The original hypothesis should be rejected.
The second hypothesis was that the surface temperature was the factor causing
oscillations. To test this hypothesis, the time of the daily temperature maximum at each
depth was plotted against time (Figure B3). The time when the linear regression line
‘Phase shift – depth” crossed the depth axis, i.e. the regression intercept, was not
significantly different from the time at which maximum capillary pressure was reached at
all depths. A similar result was obtained for water contents (data not shown). This means
that the oscillations of recorded capillary pressures and soil water contents were related to
the temperature on/near soil surface. That could be attributed to the unknown response of
the electronic equipment on the surface to the diurnal daily oscillations. To eliminate
temperature effect of measurement devices, daily time series of water content and
pressure head were averaged for whole observation period. The temporal time scale of
observations became equal to one day.
B-1
(a)
Pressure head (cm)
0
-50
-100
-150
-200
80
82
84
86
88
90
86
88
90
86
88
90
0.40
(b)
Water content (cm3/cm3)
0.38
0.36
0.34
0.32
0.30
15 cm
35 cm
55 cm
75 cm
95 cm
0.28
0.26
Temperature Co
80
82
84
22
(c)
20
18
16
14
80
82
84
Time (day)
Figure B1. Time series of capillary pressures (a), water content (b) and soil temperature
at different depths before (symbols) and after (lines) daily averaging at the location 10
cm.
B-2
Periodogram Values
40000
(a)
30000
20000
10000
0
6
12
18
24
30
36
42
48
Period (hour)
Periodogram Values
400
(b)
300
15 cm
35 cm
55 cm
75 cm
95 cm
200
100
0
6
12
18
24
30
36
42
Period (hour)
Figure B2. Periodograms for capillary pressure (a) and temperature (b) time series at
different depths.
B-3
48
Depth (cm)
0
20
40
60
12:00
06:00
12:00
06:00
Time (hour)
Figure B3. Time of the maximum temperature (∆) and maximum pressure head at 15 cm
(o) and at 35 cm (…) depth at the location “10 cm”.
B-4
Appendix C. Using temporal persistence to estimate depth-averaged
water contents with missing data
C.1 Introduction
When a field or a small watershed is repeatedly surveyed for soil water content, sites
often can be spotted where soil is consistently wetter or consistently dryer than average
across the study area. Existence of such sites is important for soil management. It is also
important for selection of sites to infer the area-average soil water content to use at
coarser scale characterization and simulation, i.e. to compare with remote sensing data or
establishing field- or catchment-wide antecedent moisture conditions for runoff
simulations (Grayson and Western, 1998). The phenomenon has been called time
stability, temporal stability, or temporal persistence in spatial patterns of soil water
content or in soil water contents. Temporal persistence of water contents at the same
depth was documented by Vachaud et al. (1985), Kachanoski and de Jong (1988), Zhang
and Berndtsson (1988), Goovaerts and Chiang (1993), Reichardt et al. (1993), and
Ferreyra et al. (2002) across areas of various extents.
Relatively less is known about temporal persistence of water content at various depths.
Comegna and Basile (1994) found a time-stable spatial structure for the water content in
the top 90 cm of the soil profile. Cassel et al. (2000) observed greater temporal
persistence of water content in deep soil layers than in shallow layers under a wheat crop.
This effect could be attributed to the impact of crop root water uptake. Hupet and
Vanclooster (2002) showed a substantial effect of vegetation on spatial and temporal
structure in profile soil water distributions. Martinez-Fernandez and Ceballos (2003) did
not observe any specific pattern of stability with respect to depth.
Our interest to the temporal persistence in soil water content in soil profile arose from the
fact that some of TDR probes positioned along soil transect at five depths with 50-cm
spacing eventually began to malfunction. Averaging data from the remaining probes at a
given depth could be an option if no persistent difference had existed between TDRmeasured water content in different locations. However such persistence clearly
manifested itself in our dataset (see, e.g., Fig. 24). Because of that, the malfunctioning of
probes with consistently lower than average values would lead to averaging only data
from the remaining probes that had water contents consistently higher than the true
average. This averaging would lead to overestimation of the average water content at the
depth of interest. In contrary, the malfunctioning of probes with consistently higher than
average values would lead to averaging only data from the remaining probes that had
water contents consistently lower than the true average, and the average water contents
would be underestimated.
To use the collected data in model calibration and evaluation, we had to propose a
technique to utilize this persistence to remedy the effect of probe malfunctioning on the
estimates of the average water content in the layer.
C.2 Temporal persistence in soil water contents
C-1
To quantify the persistence, we used an approach similar to the one proposed by Vachaud
et al. (1985). The relative water contents βij for each sampling location i at the same
depth for the measurement time j were computed as:
β ij =
θ ij
θj
(C1)
where θij is the water content measured in location i at the jth measurement time, and
θ j is the average water content at the jth measurement time at the depth of interest:
1 N
∑ θ ij
N i =1
where N is the total number of probes at a given depth (N=12 in our case).
θj =
(C2)
Statistics of values β for all probes are shown in Table C1. Of the total 60 probes, 49
probes had the values of βij that were larger or less than 1 in more than 95% of cases.
This demonstrates that temporal persistence exists in soil profile, and some probes
consistently show water contents below average whereas others show water contents
above the average. Substantial temporal linear trends in the β values with coefficients
of determination R2 larger than 0.3 were detected for four probes at the 15-cm depth, and
were absent at all other depths.
To test the dependence of the temporal stability on depth, we used values of the standard
deviations sβi of the βij values. The larger this standard deviation the less temporal
stability is observed for the probe i. Inspection of the dependence of sβi on depth showed
that there was a weak inverse relationship (data not shown). The R2 value of the linear
regression of depth vs. of sβi was 0.202 and differed statistically significantly from zero
(p=0.001).
C.3 Correcting the average water content for the persistence
Assume that K probes of total N are working at one depth. The average water content
over working probes at this depth for the sampling time j is
1 K
θ$ j = ∑ θ ij
(C3)
K i =1
The true average water content is
N
⎞
1⎛ K
θ j = ⎜ ∑ θ ij + ∑ θ ij ⎟
(C4)
N ⎝ i =1
i = K +1 ⎠
where the first term in parentheses includes data from measurements, and the second
includes unknown values that would come from malfunctioning probes should they work.
The idea of the technique is to replace the unknown values with their estimates. The
estimation consists in replacing the actual value of β ij with the average value of β ij over
the period when all probes had worked, β i
θ ij ≈ θ j β i
(C5)
C-2
where i denotes a number of malfunctioning probe (i = K+1, K+2, .., N). Values of β i are
given in Table C1.
Substituting θ ij from Eq. (C5) into Eq. (C4) and using Eq. (C3), one has
N
1⎛ ˆ
⎞
K
θ
θ
⎜
j +
j ∑ βi ⎟
N⎝
i = K +1
⎠
Rearranging Eq. (C6) leads to
Kθˆ j
θj ≈
N
N − ∑ βi
θj ≈
(C6)
(C7)
i = K +1
N
The sum of β i values is always equal to N because
N
∑θ
i =1
θj
∑ β ij =
i =1
ij
=
Nθ j
θj
= N for any j.
K
Therefore the denominator in the equation (C7) is equal to
)
following equation to estimate θ j by correcting θ j :
Kθˆ j
θj ≈ K
∑β
i =1
∑β
i =1
i
. This leads to the
(C8)
i
Only measured values are included in this equation.
To test the correction technique based on Eq. (C8), we performed simulation
experiments. We used all hourly measurements over the period when all 12 probes
worked at the 15-cm depth. First, we computed values of β i and values θ j . The latter
were dubbed ‘true average water contents.’ Then we randomly removed several probes
for each sampling time and estimated the average water content using two methods: (a)
the average water content over remaining probes assumed to be working in simulations
(i.e., θ$ j according Eq. (C3) ) , and (b) the corrected value of the average water content
according to Eq. (C8). We removed up to eight probes in the experiment I (Fig. C1a and
C1b), up to four probes in the experiment II (Fig. C1c and C1d), and one or two probes in
the experiment III. As expected, as less probes are removed, the average water content
over working probes is a better estimate of the true average (Fig. C1a, C1c, and C1e).
Correction according to Eq. (C8) consistently gave much better estimate of the true
average θ j than the average over working probes θ$ j . The improvement with Eq. (C8)
depended on the total number of working probes. The smaller the number of working
probes the more efficient was the correction.
When the correction with Eq. (C8) was applied to actual measured water contents during
the whole observation period of 360 days, it made substantial changes in some of
estimated layer-average water contents. Maximum values for the corrections of the
average layer-average water contents were 0.016, 0.013, 0.008, 0.014 and 0.014 v/v at the
depths of 15 cm, 35 cm, 55 cm, 75 cm, and 95 cm, respectively. Ranges of observed
C-3
layer-average water contents θ$ j at the same depths were 0.089, 0.062,0.037, 0.046, and
0.040 v/v, respectively. The corrections were relatively more important at larger depths
because of narrower ranges of observed water contents.
C.4 Discussion
The temporal persistence in water regimes was well-expressed at all five studied depths
(Fig. 24 and Table C1). Texture, organic matter contents, soil structure, and the number
of channels formed by roots, worms and other organisms were mentioned as leading
static factors that affect spatial variations of soil water contents (Reynolds, 1970).
Vachaud et al. (1985) related the variability of soil water contents and persistence in soil
water content distributions to the variations in soil texture. These considerations could be
pertinent to this work, although we do not have direct data to pinpoint the specific source
of the persistence. In our research area, a study of water retention along a 30-m trench in
the same soil at the adjacent site (Mallants et al., 1996) showed substantial small-scale
variability. Data on saturated volumetric water content θs from this work are shown in
Fig. 22. Values of θs were measured at 5-cm long and 5.1-cm diameter cores, i.e. at the
scale comparable with the scale of TDR measurements. The variability of saturated water
contents is comparable with variability in measured water content, indicating that the
variations in water contents measured with TDR are realistic. We hypothesized that
differences in soil structure can be responsible for those variations. The dependence of
water retention on size of soil aggregates has been demonstrated in several studies, i.e.
Wittmuss and Mazurak (1958), Tamboli et al. (1964) Amemiya (1965), Chang (1968),
and Guber et al. (2002). Such dependence can cause differences in water contents at the
same depth if soil matric potentials do not vary much across locations at this depth. This
might contribute to the temporal persistence in water contents that we observed.
The proposed approach relies on the temporal persistence. Alternatively, one could try to
estimate data for the malfunctioning probes using an interpolation technique, for example
using kriging as suggested in spatiotemporal geostatistics (Christakos, 2000). We made
an attempt to apply it, and found unsatisfactory results for probes positioned close to the
corners of the spatiotemporal domain.
The proposed correction techniques worked reasonably well, mostly because the
probability distribution functions of relative water contents were narrow, and the average
relative water content was a good approximation for the whole distribution. The wider the
distribution the less persistence is observed, and the less useful the proposed correction
can be. Temporal trends in values of relative water contents make their distributions
wider. Significant temporal trends were observed only for four probes at the 15-cm depth.
This corresponds to the conclusion of Cassel et al. (2000) about higher persistence in
deeper horizons. These authors attributed such dependence on depth to the root activity.
Plant roots were not active in this work. The weakest time persistence at the shallowest
observation depth in this work seemed to be more in line with results of Zhang and
Berndtsson (1988) and Hupet and Vanclooster (2002) who had documented weaker
temporal persistence during dry periods compared with wet ones.
C-4
Temporal persistence provides an opportunity to decrease the uncertainty of estimates of
spatially averaged water contents. Indeed, if probe readings at a given depth are thought
to be independent random values at any measurement time, the differences between those
readings will be incorporated in the standard deviation of the average water content at
this depth. However, in case of persistence, the individual probe readings are not
independent. Therefore, only the variability of relative water contents has to be included
in the estimate of the variability of the average soil water content. The variability of
relative water contents is much smaller than variability of individual probe readings.
Consequently, much smaller values of the standard deviation of the average water content
should be expected. One effect of that is an improvement in accuracy of soil water
balance computed using average soil water contents at several depths, because the
average water contents will be less uncertain (will have smaller variability estimates) if
temporal persistence in soil water content distributions is present and taken into account.
Another advantage of establishing the temporal persistence in soil water content
distributions is allowing removal some of the probes and having their measurements
reconstructed from remaining probes. The temporal persistence means a temporal
stability of spatial variability in soil water contents that may be useful in estimating
uncertainty of measurements with small number of probes. Correct treatments of those
topics are outside of the scope of this note, but they seem to be important to be explored.
In summary, the temporal persistence of water content patterns in soil profiles has been
observed at a small scale. A correction using temporal persistence has been suggested
that can be important in estimating layer-averaged water contents and their uncertainty in
case of temporary malfunctioning equipment.
C-5
0.36
a
0.32
2
R =0.864
RMSD=0.0052
0.28
0.28 0.32 0.36 0.40
0.40
c
0.36
0.32
R2=0.919
RMSD=0.0032
0.28
0.28 0.32 0.36 0.40
0.40
e
0.36
0.32
2
R =0.957
RMSD=0.0022
0.28
0.28 0.32 0.36 0.40
Corrected average soil water content over working probes, (v/v)
Average soil water content over working probes, (v/v)
0.40
0.40
b
0.36
0.32
2
R =0.988
RMSD=0.0015
0.28
0.28 0.32 0.36 0.40
0.40
d
0.36
0.32
2
R =0.996
RMSD=0.0007
0.28
0.28 0.32 0.36 0.40
0.40
f
0.36
0.32
2
R =0.998
RMSD=0.0005
0.28
0.28 0.32 0.36 0.40
True average soil water content (v/v)
Figure C1. Simulated effect of the correction for temporal persistence on the relationship
between actual and estimated layer-average water contents at the depth 15 cm; a, b –
from 1 to 8 probes removed, c, d – from 1 to 4 probes removed, e, f – from one to two
probes removed; a, b, c – average water content over remaining probes before the
correction application, d, e, f - average water content over remaining probes after the
correction application. Values of R2 are the determination coefficients, and values of the
RSMD are the root-mean square differences of the linear regression actual vs. estimated
average water contents.
C-6
Table C1. Statistics of individual TDR probes
i
dnd
βi
CVi, %
1*
131
0.93
2.0
2*
0
1.04
1.2
3
198
1.02
1.2
4*
2
0.85
1.1
5*
8
1.06
1.2
6*
0
0.94
1.9
7*
0
1.06
1.6
8
0
0.95
2.7
9*
0
0.95
1.9
10*
90
1.04
1.1
11*
132
1.09
1.4
12*
61
1.07
2.1
i
dnd
βi
CVi, %
13
114
1.01
1.1
14
0
1.00
0.9
15*
273
1.03
1.3
16*
0
0.96
0.8
17
0
0.99
0.7
18*
0
1.03
0.7
19*
0
0.89
1.1
20*
0
0.95
0.8
21*
0
0.95
1.1
22*
93
1.04
0.6
23*
212
1.07
0.6
24*
61
1.07
0.8
i
dnd
βi
CVi, %
25*
0
0.93
0.9
26
301
0.99
1.0
27*
212
0.96
0.5
28*
0
0.95
0.7
29*
0
1.02
0.7
30
0
1.00
0.7
31
0
1.02
1.5
32*
116
1.02
0.8
33
2
1.01
0.9
34*
2
1.02
0.8
35
127
1.01
2.0
36*
129
1.07
0.8
i
dnd
βi
CVi, %
37*
0
0.93
1.3
38
284
1.01
0.6
39*
0
0.97
0.6
40*
0
0.90
0.6
41*
0
1.05
0.6
42*
0
1.02
0.6
43
0
1.01
0.8
44*
166
1.07
2.4
45
109
1.01
0.6
46*
0
1.03
0.6
47*
60
1.07
0.8
48*
178
0.93
1.2
i
dnd
βi
CVi, %
49*
0
1.03
0.8
50*
254
1.05
1.3
51*
0
1.02
0.7
52*
0
1.02
0.6
53
0
1.00
0.7
54*
0
1.05
0.9
55*
0
0.84
1.3
56*
42
1.02
0.8
57
41
1.01
0.8
58
1
1.00
0.7
59*
62
1.02
0.6
60*
130
0.96
0.9
Probe number i according Fig. 1, dnd = the total number of days with no data, β i = mean
ratio of the probe “i” water contents to the depth-average water contents, CVi =coefficient
of variation of the ratios of the probe “i” water content to the depth-average water
contents. Values β i CVi were computed from hourly data. Probes marked with the
asterisk showed water contents that were different from the depth-average water contents
in more than 95% of cases.
C-7
Appendix D. Estimating matric unsaturated hydraulic conductivity in
deep soil layers
We assumed that (a) no preferential flow occurred during the drying period, (b) the
dependence of the unsaturated hydraulic conductivity on water content could be assumed
the same within the depth range from 65 to 85 cm, (c) the linear interpolation was
applicable to water contents. A simple mass balance computational scheme shown in Fig.
D1 was used.
The Darcy-Buckingham law and the mass conservation equations were used to calculate
interlayer water fluxes. Darcy equations were written for interlayer flux at depths 65 cm
and 85 cm as
⎛ h − h55
⎞
⎛ h − h75
⎞
+ 1⎟;
+ 1⎟ ,
q 65 = k 65 ⎜ 75
q85 = k 85 ⎜ 95
(D1)
⎝ 75 − 55
⎠
⎝ 95 − 75
⎠
where q is interlayer water flux, cm day-1; Kw is unsaturated hydraulic conductivity, cm
day-1; h is the capillary pressure, cm, subscripts indicate depths.
The mass conservation in the soil layer 55-75 cm is:
(D2)
∆θ75⋅20 = ∆t(q65-q85),
where ∆θ75⋅20 is daily water storage change at 20 cm layer; ∆t is time interval equal to 1
day.
Combining equation (1) and (2), one obtains:
⎛ h − h55
⎞
⎛ h − h75
⎞
(D3)
∆θ 75 ⋅ 20 / ∆t = k 65 ⎜ 75
+ 1⎟ − k 85 ⎜ 95
+ 1⎟
⎝ 20
⎠
⎝ 20
⎠
Unsaturated hydraulic conductivity values k65 and k85 are unknown variables in the
equation (D3).
We assumed the Averianov’s relationship between Kw and water content θ for depth of
65 cm and 85 cm:
n
⎛ θ −θr ⎞
⎟⎟ ,
(D4)
k = K sat ⎜⎜
⎝θs −θr ⎠
where Ksat is the saturated hydraulic conductivity; θs, θr are the saturated water content
and the and residual water content, respectively; n is exponent. The residual water
content was set at to 0.001 cm3cm-3.
Coefficients k65 and k85 in equation (3) refer to interlayer water content at 65cm and 85
cm depths, and water contents were linearly interpolated to those depths.
Equation (D3) was fitted to the water storage changes at layer 65-85 cm shown in Fig. 28.
Fitted curve described observed data adequately. The root mean square error of fitting
was equal 0.0346 cm, whereas standard deviation of measured water storage changes was
equal 0.0355 cm. Estimated values of parameters were Ksat = 0.024 cm⋅day-1, θs = 0.36
cm3cm-3, and n = 14.3.
D-1
This function was used for calculation of water flux at the depth of 105 cm as:
⎛ h − h75
⎞
+ 1⎟ − ∆θ 95 ⋅ 20 / ∆t
q105 = k 85 ⎜ 95
⎝ 20
⎠
D-2
(D5)
θ55 h55
q65
θ75 h75
q75
θ95 h95
q95
Figure D1. Water flow scheme to estimate the matric unsaturated hydraulic conductivity.
D-3
Appendix E. MWBUS: one-dimensional soil water flow model
The MWBUS (model of water balance of unsaturated soils) model has been developed to
provide a simple and robust description of water flow in layered soils with the daily time step.
The model functioning is pictured in Fig. E1.
Soil is divided in layers of thickness ∆zi, i=1,2,…, M. A downward water flow from a layer “i”can
occur if water content θi exceeds the value of water field capacity θFC,I of this layer. The value of
flux qi from the ith soil layer is calculated as:
qi = min(q1,q2)
q1= qi-1 +(θi - θFC,i)∆zi/∆t
(E1)
q2 = min(Ksat,i, Ksat,i+1)
where qi-1 is the incoming water flux to ith layer, Ksat,i is the saturated hydraulic conductivity of
ith layer, ∆t is the time interval set to one day.
The value of flux q0 is calculated from the rainfall rate R as:
q0 = min(R, Ksat,1)
(E2)
Fluxes qi, I=1,2,…,M are calculated recursively top-to-bottom. Fluxes may be corrected beginning
from any internal layer to account for complete saturation of the layer.
The layer water contents θi decrease due to evaporation and/or loss to the upward flow. The
exponential equation describes the decrease in θ with time (Suleiman & Ritchie, 2003):
θi= θr,i +(θi,initial - θr,i)exp(-t/ τ)
(E3)
where τ is the resistance parameter, θr,i is the residual water content. The daily change in water
content is
θ − θ r ,i
∆θ i = − i
τ
The value of τ increases with the depth z. Suleiman and Ritchie (2003) have found that a
power function gives a good fit to relationships between τ and z for several soils and
evaporation scenarios:
τ = τ 0 + azb
(E4)
where τ 0 is the value of resitance parameter at depth z=0, a and b are parameters of the
evaporation model.
The simulated daily evaporation rate E is:
n
E = ∑ ∆θ i ∆z i
(E6)
i =1
The number of soil layers n subject to evaporation is adjusted to make the value of E equal to
the evaporation in the model input. An adjustment over those layers is made to provide the
exact equality.
The set of the MWBUS model parameters includes the saturation water contents θs,i, the field
capacity values θFC,i, the residual water content values θr,i, and the saturated hydraulic
E-1
conductivity values Ksat,I for each layer, and the parameter of evaporation/upward loss
submodel τ0, a, and b.
a
b
Figure E1. Schematic of the modeling soil water flow with the MWBUS model; a- infiltration
regime, a downward water flow from a layer to the layer below occurs if the layer water content
exceeds the field capacity of this layer; b – evaporation regime; layers lose water exponentially in
time; the characteristic time of the water loss increases with the depth.
E-2
NRC FORM 335
U.S. NUCLEAR REGULATORY COMMISSION
1. REPORT NUMBER
(Assigned by NRC, Add Vol., Supp., Rev.,
and Addendum Numbers, if any.)
(9-2004)
NRCMD 3.7
BIBLIOGRAPHIC DATA SHEET
(See instructions on the reverse)
2. TITLE AND SUBTITLE
Model Abstraction Techniques for Soil-Water Flow and Transport
NUREG/CR-6884
3. DATE REPORT PUBLISHED
MONTH
YEAR
December
2006
4. FIN OR GRANT NUMBER
Y6724
5. AUTHOR(S)
Yakov A. Pachepsky 1, Andrey Guber 2, M.T. van Genuchten 1, T.J. Nicholson 3, R.E. Cady 3,
J. Simunek 2 and M.G. Schaap 2
6. TYPE OF REPORT
Technical
7. PERIOD COVERED (Inclusive Dates)
1 U.S. Department of Agriculture/Agricultural Research Service; 2 University of California at
Riverside; 3 U.S. Nuclear Regulatory Commission/Office of Nuclear Regulatory Research
September 2002 - August 2005
8. PERFORMING ORGANIZATION - NAME AND ADDRESS (If NRC, provide Division, Office or Region, U.S. Nuclear Regulatory Commission, and mailing address; if contractor,
provide name and mailing address.)
Environmental Microbial Safety Laboratory
Agricultural Reseach Service
U.S. Department of Agriculture
Beltsville, MD 20705
9. SPONSORING ORGANIZATION - NAME AND ADDRESS (If NRC, type "Same as above"; if contractor, provide NRC Division, Office or Region, U.S. Nuclear Regulatory Commission,
and mailing address.)
Division of Fuel, Engineering & Radiological Research
Office of Nuclear Regulatory Research
U.S. Nuclear Regulatory Commission
Washington, DC 20555
10. SUPPLEMENTARY NOTES
T.J. Nicholson, NRC Project Manager
11. ABSTRACT (200 words or less)
This report describes the methodology of model abstraction in subsurface hydrology. Model abstraction is defined as the
methodology for reducing the complexity of a simulation model while maintaining the validity of the simulation results with
respect to the question that the simulation is being used to address. The need in model abstraction may stem from the need to
improve the reliability and reduce uncertainty of simulations, to make the modeling and its results more explicable and
transparent to the end users, and to enable more efficient use of available resources in data collection and computations. A
comprehensive review of model simplification techniques developed in subsurface hydrology is included. Abstractions of both
model structure and model parameter determination are described. A systematic and objective approach to model abstraction is
outlined. A case study is presented that is designed to illustrate how model abstraction can affect performance assessment of
contaminant migration at a relatively humid site. Although the model abstraction methodology is generic, it is designed to be of
practical use to NRC licensing staff in their review of the performance assessment of decommissioning sites and waste disposal
facilities. The model abstraction process provides a systematic approach to understand the adequacy of model simplification,
and facilitates communication and transparency of the model to regulators, stakeholders, and the general public.
12. KEY WORDS/DESCRIPTORS (List words or phrases that will assist researchers in locating the report.)
flow and transport modeling
metamodeling
model abstraction
model hierarchy
model simplification
model structure
parameter estimation
pedotransfer function
soil water
subsurface hydrology
uncertainty
NRC FORM 335 (9-2004)
13. AVAILABILITY STATEMENT
unlimited
14. SECURITY CLASSIFICATION
(This Page)
unclassified
(This Report)
unclassified
15. NUMBER OF PAGES
16. PRICE
PRINTED ON RECYCLED PAPER
Printed
on recycled
paper
Federal Recycling Program
Fly UP